Вы находитесь на странице: 1из 317

Graduate School ETD Form 9

(Revised 12/07)

PURDUE UNIVERSITY

GRADUATE SCHOOL
Thesis/Dissertation Acceptance
This is to certify that the thesis/dissertation prepared
By Oscar Alfredo Ardila-Giraldo
Entitled

Investigation on the Initial Response of Beams to Blast and Fluid Impact

For the degree of Doctor of Philosophy

Is approved by the final examining committee:


Santiago Pujol
Chair

Mete A. Sozen

Ayhan Irfanoglu

Ahmed Sameh

To the best of my knowledge and as understood by the student in the Research Integrity and
Copyright Disclaimer (Graduate School Form 20), this thesis/dissertation adheres to the provisions of
Purdue Universitys Policy on Integrity in Research and the use of copyrighted material.

Santiago Pujol
Approved by Major Professor(s): ____________________________________

____________________________________
Approved by:

Rao Govindaraju
Head of the Graduate Program

27 April 2010
Date

Graduate School Form 20


(Revised 1/10)

PURDUE UNIVERSITY
GRADUATE SCHOOL
Research Integrity and Copyright Disclaimer

Title of Thesis/Dissertation:
Investigation on the Initial Response of Beams to Blast and Fluid Impact

Doctor of Philosophy
For the degree of ________________________________________________________________

I certify that in the preparation of this thesis, I have observed the provisions of Purdue University
Teaching, Research, and Outreach Policy on Research Misconduct (VIII.3.1), October 1, 2008.*
Further, I certify that this work is free of plagiarism and all materials appearing in this
thesis/dissertation have been properly quoted and attributed.
I certify that all copyrighted material incorporated into this thesis/dissertation is in compliance with
the United States copyright law and that I have received written permission from the copyright
owners for my use of their work, which is beyond the scope of the law. I agree to indemnify and save
harmless Purdue University from any and all claims that may be asserted or that may arise from any
copyright violation.

Oscar Alfredo Ardila-Giraldo


______________________________________
Printed Name and Signature of Candidate

04/23/10
______________________________________
Date (month/day/year)

*Located at http://www.purdue.edu/policies/pages/teach_res_outreach/viii_3_1.html

INVESTIGATION ON THE INITIAL RESPONSE OF BEAMS TO BLAST AND


FLUID IMPACT

A Dissertation
Submitted to the Faculty
of
Purdue University
by
Oscar Alfredo Ardila-Giraldo

In Partial Fulfillment of the


Requirements for the Degree
of
Doctor of Philosophy

May 2010
Purdue University
West Lafayette, Indiana

ii

Para Mis Padres, por su inmenso amor, incansable lucha e incondicional entrega.
Y para Caro, porque nunca te has ido; aun hoy me cuidas y sigues alumbrando mi camino.

iii

ACKNOWLEDGMENTS

This is the only section of this document that does not contain a single word
related to engineering. In spite of this (or perhaps because of this), I see this
section as the most meaningful. Here I have the opportunity to thank all those
people that one way or another have played important roles in my life and my
education. My gratitude towards them extends far beyond these pages.
I would like to express my deepest gratitude to my major advisor, Professor
Santiago Pujol, for his guidance and support, for providing the best possible
atmosphere to work side-by-side, and for always finding a way to keep up my
motivation. Learning about structural engineering from Professor Mete A. Sozen
has been an immeasurable educational experience. His contributions to this work
and his support are deeply appreciated. Special recognition and gratitude is due
to Professors Ayhan Irfanoglu and Ahmed Sameh for being part of the advisory
committee. Their valuable contributions and their kindness are deeply
appreciated. Finally, I also want to thank Professors Sozen and Sameh for their
financial support which made possible my arrival at Purdue. Without their
support, pursuing a graduate education in such a recognized institution would
have remained just a dream.
This journey through graduate school started back in Colombia when Professor
Jose D. Aristizbal-Ochoa encouraged me to pursue an education in the U.S.
Without his help, efforts, and blessings, I would have never had the opportunity
of coming to Purdue. I will be always indebted to El Profe. My deepest gratitude
also goes to Luz M. Ramirez, my peer and good friend, for her hard work which

iv
made possible that one of us was considered as a potential candidate to attend
Purdue. I was the lucky one but she deserved this opportunity as much or more
than I did.
Leaving my country and my family to come to Purdue was not easy. Fortunately,
soon after arrival at Purdue I was adopted by two of the best people I have ever
met: Jhon Paul Smith and Yeliz Firat. Their friendship made me miss my country
and my family a little bit less, and I am deeply grateful to them. My friendship with
Paul is one of the best things I got at Purdue. The same can be said about Luis
Arboleda, Fabian Consuegra, Rodrigo Segura, and David Zapata, who became
my extended family in the U.S. They are now part of a small group of people that
I regard as my all-time friends. And here I count my best friends Gonzalo
Aguilar, Ricardo Aristizbal, Juan Pablo Estrada, and Amelia Estrada. Despite
the distance, you were always part of this adventure.
Purdues ethnic and cultural diversity gave me the opportunity of meeting people
from all around the world. Some of them became very good friends of mine and
taught me that friendship is not a matter of time or common interests but genuine
connection. This was the case with Gerardo Aguilar, Carolina Gonzalez, Iva
Hrastinski, and mis nias: Vanesa Toquero y Elena Merino. Friendship with
Guillermo Cedeo was a reaffirmation that different views do not have to be a
reason for dispute but instead can motivate very enriching discussion. Heather
Akin, Rita Rodriguez, Cristina Pino, and Gustavo Cortes always had a smile for
me and their kindness and joviality was always refreshing. Carlos Arboleda and
his family were always welcoming and their home provided a space of
Colombianship more than 2000 miles away from my loved Colombia. I thank
Claudia Janssen and Becky Conley for their company and support. And I am
deeply indebted to Jillian Rumpza; without her company, awesomeness, and
selflessness, writing this thesis would have been harder.

v
My work at the Bowen Laboratory of Purdue University would have not been
possible without the constant help of Ramesh Selvarajah and Jeff Rautenberg. I
want to thank them for their patience and readiness. The assistance of Robert
Boss, Amol Godbole, Sergio Gutierrez, Kai Zhang, Lisa Choe, Harry Tidrick,
Kevin Brower, and Larry W. Skinner is greatly appreciated. Ingo Brachmann,
Seyed Hamid Schangiz, Paul Rosen, and Professor Christoph M. Hoffmann
helped me in different ways to carry out finite-element analyses using the
software LS-DYNA. And I thank Molly Stetler for her kindness, and her constant
and diligent assistance on different kinds of administrative processes.
Last but definitely not least, I want to deeply thank my family for their love,
support, and encouragement. Special thanks to my parents, Darney and Elena,
who are the real makers in this process that is now coming to an end, and to
whom I owe most of what I am. My sister, Vivi, and my brother, Ricky, have
always been essential in my life and I do not have words good enough to thank
them for their love and company. And of course I thank Caro for all her love and
goodness; she has been present all the time throughout this journey. Finally, I
want to thank some of the people in my extended family, for always being there
and believing in me: Lucecita, Gilma, my grandmother Graciela, my grandfather
Ramn, and all my aunts and uncles, but especially Elkin and his family.

vi

TABLE OF CONTENTS

Page
LIST OF TABLES ................................................................................................. x
LIST OF FIGURES .............................................................................................. xii
ABSTRACT ...................................................................................................... xxiii
CHAPTER 1. INTRODUCTION ............................................................................ 1
1.1. Background ................................................................................................ 1
1.2. Objective and Scope .................................................................................. 2
1.3. Thesis Organization .................................................................................... 3
CHAPTER 2. LITERATURE REVIEW .................................................................. 5
2.1. Introduction ................................................................................................. 5
2.2. Structural Elements Subjected to Blast Load.............................................. 5
2.3. Structural Elements Subjected to Impact Load ......................................... 13
2.4. Estimation of Maximum Dynamic Shear Demand in Beams..................... 18
2.5. Estimation of Dynamic Shear Strength in RC Beams ............................... 19
CHAPTER 3. A HYPOTHESIS ON THE OCURRENCE OF INITIAL
SHEAR FAILURE IN BEAMS SUBJECTED TO
BLAST LOAD ............................................................................... 20
3.1. Introduction ............................................................................................... 20
3.2. Tests of Aluminum Beams Loaded with Explosives ................................. 21
3.2.1. General ............................................................................................... 21
3.2.2. Experimental Plan .............................................................................. 21
3.2.3. Experimental Results .......................................................................... 22
3.3. Finite Element (FE) Simulation of Aluminum Beams Loaded
Explosively................................................................................................ 23
3.3.1. General ............................................................................................... 23
3.3.2. Finite Element Model and Simulation Details ..................................... 23
3.3.3. Finite Element Simulation Results ...................................................... 24

vii
Page
3.3.4. Discussion of Results ......................................................................... 25
3.4. Hypothesis ................................................................................................ 27
CHAPTER 4. EXPERIMENTS ON THE RESPONSE OF SMALL-SCALE
REINFORCED CONCRETE BEAMS TO FLUID IMPACT ........... 36
4.1. Introduction ............................................................................................... 36
4.2. Static Tests ............................................................................................... 37
4.2.1. General ............................................................................................... 37
4.2.2. Load-Deflection Response ................................................................. 38
4.2.3. Failure Mechanisms............................................................................ 38
4.2.4. Support Rotations ............................................................................... 39
4.3. Impact Tests ............................................................................................. 39
4.3.1. General ............................................................................................... 39
4.3.2. Measured Impact Velocities................................................................ 40
4.3.3. Response to Impact and Failure Mechanisms .................................... 41
4.3.4. Initial Damage in Impact Tests ........................................................... 50
4.3.5. Deflection Histories and Deflected Shapes......................................... 51
4.3.6. Support Rotations ............................................................................... 51
CHAPTER 5. DISCUSSION OF EXPERIMENTAL RESULTS ......................... 106
5.1. Introduction ............................................................................................. 106
5.2. Static Tests ............................................................................................. 106
5.2.1. General ............................................................................................. 106
5.2.2. Initial Stiffness and Cracking Load ................................................... 107
5.2.3. Peak Load ........................................................................................ 110
5.2.4. Maximum Shear ............................................................................... 111
5.3. Impact Tests ........................................................................................... 113
5.3.1. General ............................................................................................. 113
5.3.2. Dynamic Idealization of Specimens using SDOF Systems ............... 113
5.3.3. Idealization of Impact Load ............................................................... 116
5.3.4. Maximum Initial Shear ...................................................................... 120
5.3.5. Dynamic Shear Strength of Test Specimens .................................... 122
5.3.6. Deflections of Specimens with Flexural Mechanisms of Failure ....... 123
CHAPTER 6. ANALYSIS OF INITIAL RESPONSE OF BEAMS TO
BLAST LOAD ............................................................................. 161
6.1. Introduction ............................................................................................. 161
6.2. Characteristics of Blast Loads and Corresponding Initial
Response of Beams ............................................................................... 163

viii
Page
6.3. Classical Method to Compute Dynamic Reactions ................................. 166
6.4. Method to Compute Shear Demand using a Linear SDOF System ........ 168
6.4.1. General ............................................................................................. 168
6.4.2. Method.............................................................................................. 169
6.4.3. Variation of Factor with Time ......................................................... 172
6.4.4. Variation of Beam Stiffness with Time .............................................. 173
6.5. Computation of Shear Demand using Timoshenko Beam Theory .......... 178
6.5.1. General ............................................................................................. 178
6.5.2. Differential Equations of Motion ........................................................ 179
6.5.3. Discrete Approximation and Numerical Solution of the
Equations of Motion Using Finite Differences ................................... 181
6.5.4. Initial and Boundary Conditions ........................................................ 184
6.5.5. Implementation and Application of the Numerical Method ................ 186
6.5.6. Results for Initial Shear Demand in Beams Subjected to
Blast Load......................................................................................... 187
6.5.7. Comparison of Results from the Proposed Numerical Method
with Results from an Available Analytical Method ............................ 187
6.6. Application of the Methods Proposed to Estimate Shear Demand
to the Analysis of Initial Response of RC Slabs Tested under
Blast Load............................................................................................... 188
6.6.1. Description of Tests .......................................................................... 188
6.6.2. Load and Structure Idealizations for Analysis ................................... 189
6.6.3. Estimated Period and Shear Strength of Roof Slabs ........................ 190
6.6.4. Observed Behavior and Failure Mechanisms ................................... 190
6.6.5. Initial Deformation and Shear Demand ............................................. 191
6.6.6. Initial Shear Demand vs. Observed Behavior and Estimated
Dynamic Shear Strength................................................................... 191
CHAPTER 7. SUMMARY AND CONCLUSIONS ............................................. 217
7.1. Summary ................................................................................................ 217
7.1.1. Objective and Scope ........................................................................ 217
7.1.2. Hypothesis ........................................................................................ 218
7.1.3. Experimental Investigation of the Response of Small-Scale
Reinforced Concrete (RC) Beams to Fluid Impact ............................ 219
7.1.4. Static and Dynamic Shear Strength of Small-Scale RC Beams ....... 221
7.1.5. Response to Impact of Small-Scale RC Beams with Flexural
Mechanisms of Deformation and/or Failure ...................................... 222
7.1.6. Initial Shear Demand in Beams Subjected to Blast Load ................. 223
7.2. Conclusions ............................................................................................ 225
7.3. Future Work ............................................................................................ 228
LIST OF REFERENCES .................................................................................. 229

ix
Page
APPENDICES
Appendix A .................................................................................................... 236
Appendix B .................................................................................................... 273
Appendix C .................................................................................................... 277
Appendix D .................................................................................................... 283
VITA ................................................................................................................. 288

LIST OF TABLES

Table

Page

3.1

Aluminum Beams Subjected to Blast Load Test Results .................... 28

3.2

Mode II and Mode III Thresholds ........................................................... 29

3.3

Comparison of Experiments and Finite Element (FE) Simulations


for Beams in Series 3 ............................................................................ 30

3.4

Natural Frequencies and Periods of Vibration of Aluminum Beams ...... 30

4.1

Properties of Beams in the Different Experimental Series ..................... 52

4.2

Experimental Program Number of Specimens Tested ........................ 53

4.3

Summary of Results for Beams Tested Statically .................................. 53

4.4

Maximum Support Rotations in Static Tests Series 7-11 .................... 54

4.5

Measured Impact Velocities................................................................... 54

4.6

Midspan Deflections and Damage in Impact Tests................................ 55

4.7

Maximum Support Rotations in Impact Tests Series 7-10 .................. 56

5.1

Calculated vs. Measured Initial Stiffnesses and


Cracking Loads of Beams in the Static Tests ...................................... 136

5.2

Calculated vs. Measured Peak Loads and Maximum


Nominal Shear Stresses in Beams in the Static Tests ......................... 137

5.3

Dynamic Properties of Specimens and Equivalent


SDOF Systems .................................................................................... 138

5.4

Idealized Impact Loads and Kinetic Energies ...................................... 139

xi
Table

Page

5.5

Maximum Initial Shear in Beams in the Impact Tests .......................... 140

5.6

Initial Shear Demand vs. Failure Mechanism in the Impact Tests ....... 141

5.7

Range of Variation of Initial Shear Demand for Change in the


Failure Mode of Beams in the Impact Tests ........................................ 142

5.8

Estimated Strain Rates in the Steel Reinforcement of Beams


in the Impact Tests .............................................................................. 143

5.9

Dynamic Resistance Functions of Equivalent SDOF Systems ............ 144

5.10

Calculated vs. Measured Maximum Midspan Deflections in


Beams in the Impact Tests .................................................................. 144

5.11

Measured and Calculated (SDOF) Maximum Midspan


Deflections vs. Results from FEA - Impact Tests................................. 145

6.1

Numerical Models Used for Analysis of Initial Response of


Beams to Blast Load Variation of with Time .................................. 194

6.2

Numerical Models Used for Analysis of Initial Response of


Beams to Blast Load Maximum Shear Demand ............................... 195

6.3

Dimensions and Material Properties of RC Box Elements and


Applied Load (Slawson, 1984) ............................................................. 196

6.4

Periods of Vibration and Dynamic Shear Strengths of Roof


Slabs (Slawson, 1984) ......................................................................... 196

6.5

Estimated Initial Shear Demand vs. Dynamic Shear Strength


and Observed Damage in Roof Slabs (Slawson, 1984)....................... 197

A.1

Properties of Concrete Batches .......................................................... 249

A.2

Compressive Strength of Concrete ..................................................... 250

A.3

Modulus of Elasticity and Cylinder-Splitting Strength of Concrete ....... 252

A.4

Steel Wire Diameter Measurements Series 1-6 ............................... 252

A.5

Measured Reinforcing Steel Properties ............................................... 253

xii

LIST OF FIGURES

Figure

Page

3.1

Experimental Setup - Tests on Aluminum Beams


Loaded Explosively (Menkes and Opat, 1973) ...................................... 31

3.2

Results for Series of 6061-T6 Aluminum Beams


(0.25x1x8-in. beam) (after Menkes and Opat, 1973) ............................. 31

3.3

Stress-Strain Relationship for Material Used in FE Models ................... 32

3.4

Load Pulse Used in FE Simulations ...................................................... 32

3.5

Calculated Midspan Deflection History Tests 1-3 in Series 3 ............. 33

3.6

Calculated Beam-End Shear Force History Series 3 .......................... 34

3.7

LS-DYNA Results for the Variation with Time of End Shear (Vend)
and Deflected Shape as a Beam Subjected to Blast Load Fails
near the Supports (Test 8 in Series 3, load duration equal to
0.02 ms)................................................................................................. 35

4.1

Beam Supports ...................................................................................... 57

4.2

Schematic Description of Static Test and Drift-Ratio Definition ............. 57

4.3

Typical Views of Beams in Static Tests in Series 1, 3, 5, 7-11 .............. 58

4.4

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 1 .............................................................................. 59

4.5

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 2 .............................................................................. 59

4.6

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 3 .............................................................................. 60

xiii
Figure

Page

4.7

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 4 .............................................................................. 60

4.8

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 5 .............................................................................. 61

4.9

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 6 .............................................................................. 61

4.10

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 7 .............................................................................. 62

4.11

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 8 .............................................................................. 62

4.12

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 9 .............................................................................. 63

4.13

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 10............................................................................. 63

4.14

Applied Load vs. Beam Deflection at Midspan


Static Test_Series 11............................................................................. 64

4.15

Deflected Shape of Beam at Peak Load Static Tests_Series 1-2 ....... 65

4.16

Deflected Shape of Beam at Peak Load Static Tests_Series 3-4 ....... 65

4.17

Deflected Shape of Beam at Peak Load Static Tests_Series 5-6 ....... 65

4.18

Deflected Shape of Beam at Peak Load Static Tests_Series 7-8 ....... 66

4.19

Deflected Shape of Beam at Peak Load


Static Tests_Series 9-10........................................................................ 66

4.20

Deflected Shape of Beam at Peak Load Static Test_Series 11 .......... 66

4.21

Damage and Condition of Beam at End of Static Test Series 1 ......... 67

4.22

Damage and Condition of Beam at End of Static Test Series 2 ......... 67

4.23

Damage and Condition of Beam at End of Static Test Series 3 ......... 68

xiv
Figure

Page

4.24

Damage and Condition of Beam at End of Static Test Series 4 ......... 68

4.25

Damage and Condition of Beam at End of Static Test Series 5 ......... 69

4.26

Damage and Condition of Beam at End of Static Test Series 6 ......... 69

4.27

Damage and Condition of Beam at End of Static Test Series 7 ......... 70

4.28

Damage and Condition of Beam at End of Static Test Series 8 ......... 70

4.29

Damage and Condition of Beam at End of Static Test Series 9 ......... 71

4.30

Damage and Condition of Beam at End of Static Test Series 10 ....... 71

4.31

Damage and Condition of Beam at End of Static Test Series 11 ....... 72

4.32

Container Used in Impact Tests ............................................................ 73

4.33

Schematic Description of Impact Test ................................................... 73

4.34

Typical Views of Beams in Impact Tests ............................................... 74

4.35

Damage and Condition of Beam after Impact Specimen S1_T1 ........ 75

4.36

Damage and Condition of Beam after Impact Specimen S1_T2 ........ 75

4.37

Damage and Condition of Beam after Impact Specimen S1_T3 ........ 76

4.38

Damage and Condition of Beam after Impact Specimen S3_T1 ........ 77

4.39

Damage and Condition of Beam after Impact Specimen S3_T2 ........ 77

4.40

Damage and Condition of Beam after Impact Specimen S3_T3 ........ 78

4.41

Damage and Condition of Beam after Impact Specimen S3_T4 ........ 78

4.42

Damage and Condition of Beam after Impact Specimen S5_T1 ........ 79

4.43

Damage and Condition of Beam after Impact Specimen S5_T2 ........ 79

4.44

Damage and Condition of Beam after Impact Specimen S5_T3 ........ 80

4.45

Damage and Condition of Beam after Impact Specimen S5_T4 ........ 80

xv
Figure

Page

4.46

Damage and Condition of Beam after Impact Specimen S7_T1 ........ 81

4.47

Damage and Condition of Beam after Impact Specimen S7_T2 ........ 81

4.48

Damage and Condition of Beam after Impact Specimen S7_T3 ........ 82

4.49

Damage and Condition of Beam after Impact Specimen S7_T4 ........ 82

4.50

Damage and Condition of Beam after Impact Specimen S8_T1 ........ 83

4.51

Damage and Condition of Beam after Impact Specimen S8_T2 ........ 83

4.52

Damage and Condition of Beam after Impact Specimen S8_T3 ........ 84

4.53

Damage and Condition of Beam after Impact Specimen S8_T4 ........ 84

4.54

Damage and Condition of Beam after Impact Specimen S9_T1 ........ 85

4.55

Damage and Condition of Beam after Impact Specimen S9_T2 ........ 85

4.56

Damage and Condition of Beam after Impact Specimen S9_T3 ........ 86

4.57

Damage and Condition of Beam after Impact Specimen S9_T4 ........ 86

4.58

Damage and Condition of Beam after Impact Specimen S10_T1 ...... 87

4.59

Damage and Condition of Beam after Impact Specimen S10_T2 ...... 87

4.60

Damage and Condition of Beam after Impact Specimen S10_T3 ...... 88

4.61

Damage and Condition of Beam after Impact Specimen S10_T4 ...... 88

4.62

Damage and Condition of Beam after Impact Specimen S11_T1 ...... 89

4.63

Damage and Condition of Beam after Impact Specimen S11_T2 ...... 89

4.64

Damage and Condition of Beam after Impact Specimen S11_T3 ...... 90

4.65

Damage and Condition of Beam after Impact Specimen S11_T4 ...... 90

4.66

Cracking Patterns of Beams in Impact Tests ......................................... 91

4.67

Initial Damage in Impact Tests Specimen S1_T2 ............................... 92

xvi
Figure

Page

4.68

Initial Damage in Impact Tests Specimen S1_T3 ............................... 93

4.69

Initial Damage in Impact Tests Specimen S5_T2 ............................... 94

4.70

Initial Damage in Impact Tests Specimen S8_T2 ............................... 95

4.71

Initial Damage in Impact Tests Specimen S9_T4 ............................... 96

4.72

Initial Damage in Impact Tests Specimen S10_T3 ............................. 97

4.73

Initial Damage in Impact Tests Specimen S10_T4 ............................. 98

4.74

Initial Damage in Impact Tests Specimen S11_T3 ............................. 99

4.75

Midspan Deflection History Specimen S1_T1................................... 100

4.76

Midspan Deflection History Specimen S3_T1................................... 100

4.77

Midspan Deflection History Specimens S5_T3, S5_T4 .................... 101

4.78

Midspan Deflection History Specimens S7_T1, S7_T4 .................... 101

4.79

Midspan Deflection History Specimens S8_T1, S8_T2 .................... 102

4.80

Midspan Deflection History Specimen S9_T1................................... 102

4.81

Midspan Deflection History Specimen S10_T1................................. 103

4.82

Midspan Deflection History Specimen S11_T1................................. 103

4.83

Deflected Shape at Time of Maximum Deflection


Specimen S3_T1 ................................................................................. 104

4.84

Deflected Shape at Time of Maximum Deflection


Specimens S7_T1, S7_T4 ................................................................... 104

4.85

Deflected Shape at Time of Maximum Deflection


Specimen S8_T1 ................................................................................. 104

4.86

Deflected Shape at Time of Maximum Deflection


Specimen S9_T1 ................................................................................. 105

xvii
Figure

Page

4.87

Deflected Shape at Time of Maximum Deflection


Specimen S10_T1 ............................................................................... 105

4.88

Deflected Shape at Time of Maximum Deflection


Specimen S11_T1 ............................................................................... 105

5.1

Shear Force and Moment Diagrams for Fixed-Fixed Beam


with Concentrated Load at Midspan .................................................... 146

5.2

Calculated Moment-Curvature Relationships for Beams


in Series 1-2 ........................................................................................ 147

5.3

Calculated Moment-Curvature Relationships for Beams


in Series 3-4 ........................................................................................ 147

5.4

Calculated Moment-Curvature Relationships for Beams


in Series 5-6 ........................................................................................ 148

5.5

Calculated Moment-Curvature Relationships for Beams


in Series 7-8 ........................................................................................ 148

5.6

Calculated Moment-Curvature Relationships for Beams


in Series 9-11 ...................................................................................... 149

5.7

Representation of Test Specimen with an Equivalent


SDOF System ...................................................................................... 149

5.8

Elasto-Plastic Resistance Function of Equivalent SDOF System ........ 150

5.9

Idealized Impact Load.......................................................................... 150

5.10

Initial Impact Loading and Beam Response ........................................ 151

5.11

Variation of Initial Shear Demand for Change in the Failure


Mode and Suggested Dynamic Shear Strength of Beams in
the Impact Tests .................................................................................. 151

5.12

Calculated Midspan Deflection History Specimen S1_T1 ................. 152

5.13

Calculated Midspan Deflection History Specimen S3_T1 ................. 152

5.14

Calculated Midspan Deflection History Specimen S5_T3 ................. 153

xviii
Figure

Page

5.15

Calculated Midspan Deflection History Specimen S5_T4 ................. 153

5.16

Calculated Midspan Deflection History Specimen S7_T1 ................. 154

5.17

Calculated Midspan Deflection History Specimen S7_T4 ................. 154

5.18

Calculated Midspan Deflection History Specimen S8_T1 ................. 155

5.19

Calculated Midspan Deflection History Specimen S9_T1 ................. 155

5.20

Calculated Midspan Deflection History Specimen S10_T1 ............... 156

5.21

Calculated Midspan Deflection History Specimen S11_T1 ............... 156

5.22

Typical Finite Element Model to Calculate Response of


Small-Scale RC Beams to Fluid Impact using LS-DYNA ..................... 157

5.23

Typical Rendering of LS-DYNA Simulation of Fluid Impact on


Small-Scale RC Beams (Specimen S3_T1, 7.8 ms after impact) ........ 157

5.24

Calculated (FEA) Midspan Deflection History Specimen S3_T1 ...... 158

5.25

Calculated (FEA) Midspan Deflection History Specimen S7_T1 ...... 158

5.26

Calculated (FEA) Midspan Deflection History Specimen S8_T1 ...... 159

5.27

Calculated (FEA) History of Strain in the Reinforcement


in Tension Specimen S3_T1............................................................. 159

5.28

Calculated (FEA) History of Strain in the Reinforcement


in Tension Specimen S7_T1............................................................. 160

5.29

Calculated (FEA) History of Strain in the Reinforcement


in Tension Specimen S8_T1............................................................. 160

6.1

Pressure-Time History for Blast Wave ................................................. 198

6.2

Determination of Dynamic Reactions (Biggs, 1964) ............................ 199

6.3

Static Deflected Shape vs. Initial Deformed Shape under


Blast Load - Beam with Fixed Ends ..................................................... 200

xix
Figure

Page

6.4

Dynamic Reactions in the Initial Phase of Response to


Blast Load............................................................................................ 201

6.5

Hypothetical Initial Deformed Shape of a Beam under Blast Load ...... 202

6.6

Deflected Shapes of Fixed-Fixed Beams at Time of Maximum


Shear in the Initial Phase Calculated Using LS-DYNA
Beams 1, 3, 5 ...................................................................................... 202

6.7

Deflected Shapes of Fixed-Fixed Beams at Time of Maximum


Shear in the Initial Phase Calculated Using LS-DYNA
Beams 2, 4, 6 ...................................................................................... 202

6.8

Deflected Shapes of Fixed-Fixed Beams at Time of Maximum


Shear in the Initial Phase Calculated Using LS-DYNA
Beams 7-10 ......................................................................................... 203

6.9

Deflected Shapes of Fixed-Fixed Beams at Time of Maximum


Shear in the Initial Phase Calculated Using LS-DYNA
Beams 11-14 ....................................................................................... 203

6.10

Deflected Shapes of Fixed-Fixed Beams at Time of Maximum


Shear in the Initial Phase Calculated Using LS-DYNA
Beams 15-16 ....................................................................................... 203

6.11

Variation of with Time in the Initial Phase for Numerical


Models of Fixed-Fixed Beams Subjected to Blast Load ...................... 204

6.12

Idealized Beam Deformation and Stiffness in the Initial


Phase of Response to Blast Load ....................................................... 205

6.13

Deformations and Dynamic Equilibrium of a Beam


According to the Timoshenko Theory .................................................. 206

6.14

Approximation of the Two-Dimensional Space-Time Domain


of the Equations of Motion of a Beam According to the
Timoshenko Theory and Rectangular Grid Used for
Numerical Solution Using Finite Differences........................................ 207

6.15

Deflected Shapes of Fixed-Fixed Beams at Time of Maximum


Shear in the Initial Phase Calculated Using TBT and LS-DYNA.......... 208

xx
Figure

Page

6.16

History of Shear and Moment at the Support of Slab Subjected to


Blast Load (after Ross and Krawinkler, 1985) ..................................... 209

6.17

History of Shear and Moment at the Support of Slab Subjected to


Blast Load (computed using TBT and algorithm in Appendix D) ......... 210

6.18

RC Box Elements Tested under Blast Load by Slawson (1984) .......... 211

6.19

Typical Posttest View of RC Box Element in Series FY81


Concrete Crushing and Reinforcement Pullout at the Supports
of the Roof Slab (after Slawson, 1984) ................................................ 212

6.20

Idealization of Roof Slabs of RC Box Elements ................................... 212

6.21

Typical Posttest Views of RC Box Elements in Series FY82


Deformation/Failure Dominated by Shear (after Slawson, 1984)......... 213

6.22

Calculated Maximum Initial Shear Demand in Roof Slabs................... 214

6.23

Deflected Shapes at Times of Maximum Initial Shear Resisted


by Roof Slabs Calculated using TBT ................................................... 215

6.24

Deflected Shapes at Times of Maximum Initial Shear Resisted


by Roof Slabs Calculated using TBT ................................................... 215

6.25

Shear Strength of RC Joints Subjected to Pure Shear vs.


Reinforcement Ratio (after Birkeland, 1966, and Phillips, 1972) ......... 216

A.1

Compressive Strength vs. Time - Concrete Batches 1 and 3 .............. 254

A.2

Compressive Strength vs. Time - Concrete Batches 4-6..................... 254

A.3

Compressive Strength vs. Time - Concrete Batches 7-8..................... 255

A.4

Compressive Strength vs. Time - Concrete Batches 9-11................... 255

A.5

Initial Stress-Strain Curves for Concrete Cylinders Batch 7 ............. 256

A.6

Initial Stress-Strain Curves for Concrete Cylinders Batch 8 ............. 256

A.7

Initial Stress-Strain Curves for Concrete Cylinders Batch 9 ............. 257

A.8

Initial Stress-Strain Curves for Concrete Cylinders Batch 10 ........... 257

xxi
Figure

Page

A.9

Initial Stress-Strain Curves for Concrete Cylinders Batch 11 ........... 258

A.10

Concrete Modulus of Elasticity vs. Compressive Strength .................. 259

A.11

Concrete Tensile (Cylinder Splitting) vs. Compressive Strength ......... 259

A.12

Representative Stress-Strain Curves for Steel Wires


Series 1-6 ........................................................................................... 260

A.13

Representative Stress-Strain Curves for Steel Wires Series 7-11 ......................................................................................... 260

A.14

Test Specimen: Nominal Dimensions and Reinforcement Details ...... 261

A.15

Molds Prepared for Casting................................................................. 262

A.16

Placing and Compaction of Concrete .................................................. 262

A.17

Specimen Surfaces after Being Leveled and Smoothed ..................... 263

A.18

Curing of Concrete .............................................................................. 263

A.19

Specimen Supports and Reaction Structure


(all dimensions in inches) .................................................................... 264

A.20

Static Test Setup ................................................................................. 265

A.21

Impact Test Setup ............................................................................... 266

A.22

Location of Deflection Measurements ................................................. 267

A.23

Protection of LVDTs against Impact and Attachment to


Specimen............................................................................................ 268

A.24

Protection and Attachment of LVDTs at Upper Support Plate ............. 268

A.25

Typical Displacement Histories of Upper Support Plate in


Series 11 ............................................................................................ 269

A.26

Typical Displacement Histories of Upper Support Plate in


Series 10 ............................................................................................ 269

xxii
Figure

Page

A.27

Photodetector Signal Records for Measurement of


Impact Velocity Series 1_Test 1 ...................................................... 270

A.28

Photodetector Signal Records for Measurement of


Impact Velocity Series 8_Test 2 ...................................................... 270

A.29

Position of High-Speed Camera in Impact Tests................................. 271

A.30
A.31

Foam-Board Flaps on Both Sides of Specimen in Impact Tests ....... 271


Empirical Relationship between Pressure of Helium Used as
Propellant and Outcome Velocity of Projectile .................................... 272

B.1

Calculated Displacement History of Equivalent SDOF System


(Specimen S8_T1) ............................................................................... 276

B.2

Resistance-Displacement Relationship for Equivalent


SDOF System (Specimen S8_T1) ...................................................... 276

C.1

Pressure vs. Time Curve for Blast Load


(Beam ID = 6 in Table 6.2) .................................................................. 280

C.2

Total Load applied to the Beam vs. Time ............................................ 280

C.3

Displacement History of Equivalent Linear SDOF System .................. 281

C.4

Variation of Factor with Time ........................................................... 281

C.5

Maximum Shear (Reaction Force) in the Initial Phase ........................ 282

xxiii

ABSTRACT

Ardila-Giraldo, Oscar Alfredo Ph.D., Purdue University, May 2010. Investigation


on the Initial Response of Beams to Blast and Fluid Impact.
Major Professor: Santiago Pujol.

Experimental evidence shows that an element that can develop a flexural


mechanism of failure under static load may fail in shear under blast load. This
study shows that the critical parameter defining the failure mode of an element
under blast or impact load is the ratio of shear demand to shear capacity. The
matter is not as simple as it may appear because for dynamic load, shear
demand depends on inertial forces and shear capacity depends on loading rate.
This study provides means to estimate dynamic shear demand.
The ratio of shear demand to shear capacity controlled the failure mechanism of
small-scale reinforced concrete (RC) beams tested under fluid impact. The test
specimens had their ends restrained against rotation, and were subjected to
transverse impact at midspan of liquid bodies traveling at speeds of up to 150
m/s. The specimens were observed to fail in shear at nominal shear stresses
above approximately 1.6 times their static shear strengths. The threshold for
shear failure seemed independent of vibration properties. Control specimens
developed flexural mechanisms of failure under static load. The tests showed
that the change in failure mode observed in aluminum beams subjected to
distributed blast load (Menkes and Opat, 1973) can also occur in small-scale RC
beams subjected to concentrated fluid impact.

xxiv
The deformed shape of a beam in the initial phase of response to blast load (0 t
0.1T, where T is the vibration period) is dominated by shear deformations near
the supports. Deformations progress from the supports towards midspan as the
beam acquires a shape similar to the deflected shape under static load at a time
longer than 0.1T. The initial deformed shape is dramatically different from the
deformed shape under static load. The deformed shape affects shear demand.
Two numerical methods to estimate initial shear demand caused by blast load
are proposed. The methods account for the mentioned change in the deflected
shape. One method is based on the work by Biggs (1964) to compute dynamic
reactions, and uses a single-degree-of-freedom (SDOF) system with timedependent stiffness. The other method involves numerical solution of the
differential equations of motion of a beam according to the theory proposed by
Timoshenko (1937), which includes the effects of shear deformations and
rotational inertia. Both methods assume the response of the beam to be linear
and provide estimates of shear demand that are consistent with the experimental
observations available to date.

CHAPTER 1. INTRODUCTION

1.1. Background
Blast or high-speed impact can cause disproportionate structural damage
following local damage. One measure to prevent disproportionate damage is to
design critical structural members to resist the effects of extreme loads. The
objective is to design the critical members to be tough and ductile so that their
energy-dissipation capacity is high. These elements are assumed to dissipate
energy through flexural deformation. It is desirable that the elements be able to
reach large deflections without failing in shear. To ensure this condition,
engineers typically: (1) design the elements to resist shear forces equal to or
larger than those associated with a static flexural mechanism of failure, and (2)
detail the elements to accommodate the large rotations and ductility demands
associated with a flexural mechanism.
Experimental evidence (Menkes and Opat, 1973) shows that a structural element
that can develop a flexural mechanism of failure under static load may fail in
shear under dynamic load of high intensity and short duration. In this case, the
ability of the element to dissipate energy is limited by shear failure. Because
energy-dissipation capacity is the key to the survival of the element to blast or
impact load, shear failure must be avoided. To do so, the likelihood of a change
in the failure mode (from flexure to shear) of the element under blast or impact
load must be quantified.
The phenomena related to the described change in failure mode of a beam under
dynamic load have been studied by others before. Chapter 2 contains a detailed

2
review of previous work on the subject. This review reveals lack of consensus
about the critical parameters affecting the change in failure mode. It also shows
that the analytical methods currently used to describe the problem are complex
and deal mostly with metal beams. These limitations of the state of the art form
the motivation of the study presented here, which focuses on the need for
simpler design tools and reinforced concrete (RC) elements.

1.2. Objective and Scope


The objective of the study described in this report is to obtain a simple way to
determine whether a beam that fails in flexure under static load may fail in shear
under blast or fluid impact.
Methods to determine the likelihood of shear failure under dynamic load are
developed from a simple perspective. It is suggested that the critical parameter is
the ratio of shear demand to shear capacity. For dynamic load, estimation of
shear demand requires consideration of inertial forces. This is done in this study
through numerical analysis (chapter 6). The capacity of an element to resist
shear is sensitive to loading rate. The effect of loading rate on shear capacity
was studied experimentally for small RC beams (chapter 4 and chapter 5).
The scope of this study included (a) construction, instrumentation, and testing of
forty-two small-scale RC beams under static load and impact of a liquid body,
and (b) implementation of numerical methods to compute initial shear demand on
a beam subjected to dynamic load.
This study does not consider axial load acting on a structural element subjected
to blast or impact load. The study is focused on beams with:
(1) End rotations fully restrained,
(2) loads perpendicular to the longitudinal axis, and

3
(3) static shear strengths higher than shear forces at yield (i.e., beams that
can develop their static flexural capacities without failing in shear).
Tests were done to investigate whether the phenomenon observed by Menkes
and Opat (1973) can also occur for RC beams and for concentrated dynamic
loads. The experimental program and its results (chapter 4) are limited to:
(1) Small-scale RC beams 1-in. deep, with clear spans of 8 in. and 12 in., and
widths ranging from 2 in. to 6 in.
(2) Beams with continuous top and bottom longitudinal steel reinforcement
and without transverse reinforcement.
(3) Longitudinal reinforcement ratios () of 0.5% and 1% (identical for top
and bottom reinforcement), and longitudinal reinforcement yield stresses
(fy) of 50 ksi and 120 ksi.
(4) Compressive strength of concrete (fc) in the range 6-10 ksi.
(5) Ratio of shear span (a) to effective depth (d) equal to 2.7 and 4.
(6) Concentrated load (static and fluid impact) at midspan.
(7) Impact velocity (vp) ranging approximately from 50 m/s to 150 m/s.
(8) Ratio of duration of impact load (td) to calculated period of vibration of the
impacted beam (T) ranging approximately from 0.4 to 1.4.
(9) Ratio of mass of impacting liquid body (m) to effective mass of the
impacted beam (Me) ranging approximately from 0.2 to 0.5.

1.3. Thesis Organization


Chapter 2 reviews relevant existing research on the response of structural
elements to blast and impact load. Methods to estimate shear demand and
dynamic shear strength of elements under dynamic and impulsive load are
reviewed as well.

4
Chapter 3 presents the experimental results obtained by Menkes and Opat
(1973), which motivated the study described in this report. These results point
towards a simple hypothesis that is presented at the end of the chapter.
Chapter 4 summarizes the results of the experimental program conducted on
small-scale RC beams subjected to concentrated dynamic (fluid impact) load.
These results show that what Menkes and Opat (1973) observed for aluminum
beams subjected to uniform blast load can also occur for RC beams and for
concentrated dynamic loads.
Chapter 5 provides a discussion of the experimental results (chapter 4). A
relationship between initial shear demand and failure mechanism is presented for
the specimens tested.
Chapter 6 deals with the initial response of beams to rapidly applied uniform load
(blast load). Two methods are proposed to estimate initial shear demand on
beams subjected to blast load.
Chapter 7 presents summary and conclusions of this study, as well as
recommendations for future work.

CHAPTER 2. LITERATURE REVIEW

2.1. Introduction
A beam/column under blast or impact load may experience large plastic
deformations and/or fail. The key to its survival is its energy-dissipation capacity.
Newmark (1953) suggested that a structural element could be designed to resist
an impulsive load (blast or impact) simply by ensuring that the energy-dissipation
capacity of the element is equal to or larger than the energy transferred by the
load.
Experimental data show that beams that fail in flexure under static load may fail
in shear under blast load (Menkes and Opat, 1973). This change in failure
mechanism (from flexure to shear) affects the ability of an element to absorb the
energy imparted by the load. A summary of existing experimental and analytical
research on the response, modes of deformation, and failure mechanisms of
structural elements under blast and impact load is presented in this chapter.
Chapter 3 suggests that the problem of shear failure of elements under blast or
impact load can be studied simply by comparing the maximum shear demand
with the capacity of the element to resist shear forces. Available literature on the
estimation of maximum shear demand and dynamic shear strength of elements
loaded rapidly is also reviewed in this chapter.

2.2. Structural Elements Subjected to Blast Load


Symonds and Mentel (1958) used rigid-plastic beam theory to study the nonlinear response of a beam with axial constraints to uniform transverse impulsive

6
load. A two-dimensional yield condition with interaction between bending moment
and axial force was used. Assuming infinitesimal displacements, a solution for
the maximum permanent deflection of the beam was obtained. For the range of
response in which the beam behaves as a plastic string, lower and upper
bounds on the final plastic deformation of the beam were obtained. The effects of
shear forces and deformations on the response of the beam were not addressed.
The response of a beam with axial constraints to uniform transverse impulsive
load was studied experimentally by Humphreys (1965). He focused on the
response of straight rectangular steel beams fully restrained axially (fixed ends)
and loaded with explosives. Different beam sizes and material properties were
used in the tests. Blast loads were large enough to produce plastic deformations.
The observed permanent deflections of the tested beams were compared with
the upper bounds given by Symonds and Mentel (1958). These upper bounds
were in general higher (by approximately 20-30%) than the deformations
observed experimentally. It was concluded that a good approximation to the
maximum deflection of a beam subjected to blast load is obtained simply by
using rigid-plastic beam theory (including axial constraints). A shear failure mode
was not observed in any of the beams tested.
Menkes and Opat (1973) did a series of experiments on 6061-T6 aluminum
beams with fixed ends and loaded with explosives. Different beam geometries
were considered in the tests. Similar beams showed different failure mechanisms
depending on the intensity of the impulsive load. As the load was increased, the
failure mode changed from a flexural mechanism to shear failure at the supports.
At intermediate loading levels, a failure mode referred to as tensile tearing was
observed (rupture of outer fibers in tension at the ends of the beams). Beams
that failed in shear at the supports showed nearly zero curvature in the central
portion of their spans. The test results were used to suggest that the failure mode
of a beam is primarily determined by the applied impulse per unit loaded area or,
equivalently, the initial velocity imparted to the beam. In these tests, shear failure

7
was observed for specimens loaded to have an initial average velocity exceeding
10,000 in/s (~830 ft/s). For a given beam cross section, the smallest unit impulse
that caused shear failure at the support was independent of beam length.
Jones (1976) used rigid-plastic beam theory to estimate the permanent mid-span
deflection and the threshold impulsive velocity at which shear failure at the
support occurs for a metal beam with fixed ends and subjected to blast load. The
results of these approximate theoretical methods were in good agreement with
the experimental results reported by Menkes and Opat (1973). A previously
developed theoretical procedure that considers finite deflections in the dynamic
plastic response of beams (Jones, 1971) was used to predict permanent
deflections. The threshold velocity at which shear failure occurs was estimated
following a procedure similar to the one proposed by Symonds (1968) to study
the influence of shear forces and deformations on the behavior of rigid-plastic
beams loaded impulsively. It was assumed that the central portion of the beam
moved with a velocity equal to the impulsive velocity applied by the blast load,
and that the ends of the beam moved at a lower velocity as a result of the lateral
restraint provided by the supports. Shear sliding was assumed to take place at
the end sections of the beam. It was assumed that complete severance occurs
when the sliding at the end section equals the thickness of the beam. By
enforcing this condition, and considering conservation of momentum and rigidplastic response, a threshold velocity for shear failure at the support of a beam
subjected to blast load was obtained.
Quarter-scale reinforced-concrete (RC) box-type structures were tested under
blast load at the US Army Engineer Waterways Experiment Station, WES
(Slawson, 1984). The box elements had flat roof slabs, and were buried and
tested in a sand backfill. The purpose of these experiments was to investigate
the possibility of shear failure in shallow-buried RC box elements subjected to
nuclear blast. Airblast simulation techniques involving chemical explosives were

8
used to simulate the blast load resulting from a nuclear detonation. The
parameters varied in the experiments were concrete strength, structural stiffness,
steel reinforcement ratio, and charge density. Deformations and/or failure modes
of the roof slabs were recorded using high-speed video. The slabs were
observed to deform or fail in flexure or shear depending on the intensity of the
blast load. Slabs subjected to larger blast loads showed damage attributable to
shear and were either completely severed from the walls of the box elements
(along failure planes parallel to these walls), or had concentration of deformation
near their ends and their central portion was relatively flat.
An analytical method to investigate the dynamic shear failure of RC beams and
one-way slabs under blast load was proposed by Ross and Krawinkler (1985).
The effects of load rate and beam-end restraint were investigated. A linear
Timoshenko beam subjected to a rapidly applied uniform load, the magnitude of
which decreased linearly with time, was analyzed using the normal mode
method. For a given set of beam and load parameters (beam geometry,
stiffness, and strength, rise time, load duration, and peak pressure), solutions of
the Timoshenko beam equations for bending moment and shear, normalized with
respect to the estimated moment and shear capacities of the beam, respectively,
were obtained. From these solutions, the variation with time of (normalized)
moment and shear was obtained. It was observed that at early times, shear
increases at a higher rate than moment, and at later times, moment increases
faster than shear. The times at which the calculated normalized moment and
shear equal 1 were determined. A beam was assumed to fail in direct shear if the
time required for the normalized shear to equal 1 is shorter than the
corresponding time for the normalized moment. If both times are equal, a
transition in the failure mode was assumed to occur. Transitions for a given
beam subjected to different load curves were obtained from different analyses.
These transitions were plotted in the peak-pressure vs. rise-time diagram and a
failure curve for the beam was obtained. This failure curve was used to predict

9
whether a beam subjected to a given load pulse would fail in shear or bending.
The developed shear-failure curves were used to analyze the behavior of the roof
slabs of the RC box elements tested under blast load by Slawson (1984).
Shen and Jones (1992) proposed an energy criterion to predict the inelastic
failure modes of a beam subjected to blast load. The most relevant parameter
considered was the ratio of plastic shear work to the total plastic work dissipated
at regions of the beam where large plastic strains are expected to concentrate
(near supports or points of load application). The variation of this parameter was
obtained by analyzing theoretically the response of a rigid-perfectly plastic beam
with fixed ends to a uniform impulsive load. Plastic yielding of the beam was
assumed to be governed by an interaction yield surface that combined bending,
shear, and tension effects. The influence of material strain-rate sensitivity in
beam response was studied using the relationship proposed by Cowper and
Symonds (1957). The failure modes observed by Menkes and Opat (1973) were
examined theoretically using the proposed energy criterion. It was suggested that
the dynamic inelastic failure of beams occurs when the plastic work density
reaches a critical value.
Wen, Yu, and Reddy (1995) used a quasi-static procedure to obtain the dynamic
plastic response and failure of clamped beams subjected to uniform impulsive
load. The quasi-static procedure used was an adaptation of the procedure
previously developed by Wen, Reddy, and Reid (1995) to estimate the response
of clamped beams to impact load. The principle of conservation of energy and
the concept of equivalent work were used to analyze a beam under impulsive
load using the simpler case of an identical beam subjected to an equivalent static
load. A transition curve from tensile tearing failure to shear failure of the beam
was obtained. It was assumed that tensile tearing and shear failure occur when
the rupture strain of the beam material and the critical shear force, respectively,
are reached. The predicted values of permanent beam deflections and critical

10
impulses that cause tensile tearing and shear failure were in good agreement
with the test data reported by Menkes and Opat (1973). It was shown that the
onset of the failure modes mentioned before depends on the mechanical
properties of the material, beam geometry, and support conditions. Similar
conclusions had been obtained by Wen, Reddy, and Reid (1995) for beams
subjected to impact load.
A bomb attack took place at the Alfred P. Murrah Federal Building (referred to
here as the Murrah Building) in Oklahoma City, Oklahoma, on 19 April 1995.
Damage observed in the Murrah Building was reported in detail by a building
performance assessment team (BPAT) deployed by the Federal Emergency
Management Agency (FEMA, 1996). Major structural damage and partial
collapse occurred near the north face of the Murrah Building (facing the
explosion), which was an ordinary RC-frame structure. The BPAT determined
that the major structural damage caused directly by the blast was the loss of
three exterior columns on the north face of the building, as well as portions of
some floors. The three exterior columns supported a transfer girder at the thirdfloor level which provided support to intermediate columns above the third floor.
The loss of exterior columns below the third-floor level left the transfer girder
unsupported, triggering the progressive collapse of floors above. According to the
BPAT, the exterior column that was closest to the explosive charge was removed
abruptly by the blast through brisance (a shattering effect). The other two exterior
columns destroyed by the blast failed in shear before developing flexural
mechanisms. Loading on these columns caused near-yield moments to develop
over the heights of these elements, and the corresponding shear stresses at the
end sections were estimated to have exceeded the calculated shear capacities of
the columns.
Woodson and Baylot (1999) tested five RC frames under blast load at the US
Army Engineer WES. These tests were conducted to investigate the response of

11
RC building columns to blast load, including the effects of the presence of infill
masonry walls. The test specimens were quarter-scale flat-plate building models.
They were two-stories tall, two-bays wide, and one-bay deep. The four corner
columns of each specimen were not regarded as test columns, and they were
oversized and provided with high reinforcement ratios. The two center columns
were considered test columns, and they represented an exterior and an interior
column in an RC flat-plate building. Blast pressure records and photographs of
deflected shapes of the test columns were obtained. In three out of four
experiments, the exterior test column showed concentration of deformation and
severe damage (apparently caused by shear) near its ends, with the central
portion of its length remaining flat. In the remaining experiment, which had a
lower blast load, the exterior test column was not damaged.
Li and Jones (1999) revisited the problem of shear failure at the support of a
fully-clamped metal beam subjected to impulsive pressure. Following the work by
Jones (1976), rigid-perfectly plastic response was assumed for the portion of the
beam between the shear hinges developed at both ends of the span. Within the
shear hinges (where shear sliding takes place), strain hardening and strain rate
effects were included to represent better the failure mechanism. It was assumed
that shear failure occurs when the sliding at the end section equals a certain
fraction of the thickness of the element. An expression for the critical impulsive
velocity causing shear failure at the support of the beam was obtained.
Yu and Chen (2000) also studied the shear failure at the support of a clamped
metal beam loaded impulsively. The failure criterion adopted by Jones (1976)
was modified to include the weakening of sliding sections (beam end sections)
during failure. Considering the variation with time of shear and bending moment
at the end sections, the plastic deformation and failure mechanism of the beam
were traced, and the ratio of plastic shear work to total plastic work dissipated
was calculated. Interaction between shear and bending at the sliding sections

12
was considered by using three different yield curves. Complete severance at the
supports was assumed to occur when the sliding at the end section is k (k < 1)
times the thickness of the beam. The factor k was thought to be larger for a tough
material than for a brittle material. It was indicated that k can be approximated as
a constant (k = 0.3 was suggested) for a wide range of situations. A shear-strain
failure criterion was also discussed. An expression for the threshold impulsive
velocity that causes shear failure at the support was obtained. Predictions of this
threshold velocity were compared with results from other theoretical approaches
and with experimental data (Menkes and Opat, 1973).
Relevant existing research on response of metal structures to blast load was
reviewed by Zhu and Lu (2007). The focus of this work is the characteristics of
blast load and corresponding structural response, as well as the theoretical and
experimental advances in this area. The concept of blast wave and its effects on
a structure were discussed. The main characteristics of the response of structural
elements to blast load were identified as (1) mode of deformation and fracture,
(2) impulse transfer, and (3) energy absorption in plastic deformation. The
major experimental methods, devices, and sensors used in the investigation of
blast loading and measurement of load parameters such as impulse or pressure,
or structural response parameters such as acceleration and displacement were
summarized. Relevant experimental studies that have shown different modes of
deformation and failure of metal structures subjected to blast load were listed.
The most common analytical models used to determine the dynamic response of
a structure to blast load were reviewed. This included (1) single-degree-offreedom (SDOF) models, (2) modal approximation models, and (3) models based
on rigid plasticity theory. Approaches to numerical modeling of blast loads and
dynamic behavior of materials for use in finite element analysis (FEA) to evaluate
the response of metal structures to blast load were discussed as well.

13
2.3. Structural Elements Subjected to Impact Load
One of the first formulations to estimate the force exerted on a flat solid barrier by
an impacting aircraft was proposed by Riera (1968). This work is applicable to
normal impact on a rigid target. The proposed expression for impact force was
derived by considering dynamic equilibrium (conservation of linear momentum) of
the impacting body. Using the impact velocity of the aircraft as well as its mass
and strength distribution, the force-time relationship for the impact of a typical
commercial aircraft on a rigid wall was obtained. A similar result was obtained
experimentally by Sugano et al. (1993), who conducted a full-scale aircraft
impact test to evaluate the force exerted on a massive concrete block. Response
measurements obtained during the test were used to determine the impact forcetime curve, which was in good agreement with the curve obtained analytically
using conservation of linear momentum (Riera, 1968).
Liu and Jones (1987) did a series of tests on clamped metal beams subjected to
impact load. Both aluminum and steel beams were used in the tests. Different
beam thicknesses were considered. A drop mass apparatus was used to apply
the dynamic load. The impact point was varied from mid-span to the immediate
vicinity of a support. Different drop heights for the impacting mass were used to
examine the influence of impact velocity on the response of the beams. The
external energy imparted to the beams had a wide range of variation so that
beam responses varying from small permanent deflections to shear failures could
be attained. Strain gages were attached to both the upper and lower surfaces of
more than 10 beams to obtain indirect measurements of bending moments. Highspeed videos were recorded. In several tests, beams underwent bending and
had permanent deflections. In other tests, beams failed in two different modes,
classified as tensile tearing and shear failure. The threshold impact velocities
for these failure modes seemed to depend on the rupture strain of the material
and the location of the impact point. An important decrease in energy-absorption
capacity was observed for beams impacted near a support.

14
Krauthammer et al. (1990) proposed a numerical procedure for the analysis of
RC beams and one-way slabs subjected to impulsive load. Special attention was
given to the response of RC beams to localized dynamic loads such as impact
load. Based on experimental observations of RC slabs under blast load that
failed in direct shear at the supports (Kiger and Slawson, 1984; Slawson, 1984)
before significant flexural response was developed, Krauthammer et al. (1990)
concluded that it was reasonable to decouple the direct shear response from the
flexural response of a beam loaded impulsively. A model with two separate,
nonlinear SDOF systems to evaluate the flexural and direct-shear responses of a
beam to impulsive load was proposed. Beam resistance functions corresponding
to the two different types of response examined were developed. The dynamic
analysis method proposed used variable SDOF parameters (equivalent mass,
load, and stiffness), which were computed at every time step. This required the
computation of the deformed shape of the beam at every time step as well. Thus,
dynamic reactions were obtained using an instantaneous deformed configuration
of the beam. Linear beam theory was used in the computation of the dynamic
reactions, based on the fact that peak direct shear response always occurs early
in time, before any significant nonlinear flexural response is developed. Using
the SDOF system for evaluation of flexural response, the deflections and
dynamic reactions of the beam were obtained. These reactions were applied as
forcing functions to the SDOF system for evaluation of shear response. At every
time step, the computed shear response was compared with a failure criterion for
direct shear. A direct-shear failure was assumed to occur when the shear slip at
the critical section exceeds the ultimate permissible shear displacement of the
beam (Krauthammer et al., 1986).
A commercial airliner was crashed into the Pentagon Building in Arlington,
Virginia, on September 11, 2001. The American Society of Civil Engineers
(ASCE) established a building performance study (BPS) team that inspected the
Pentagon Building site and investigated the damage to the structure (Mlakar et

15
al., 2003). One of the aspects analyzed by the BPS team was the response of
the structural system of the Pentagon to the impact of the airplane. It was
concluded that approximately 50 columns were destroyed or heavily damaged by
the impact. Impact analyses showed that the spirally reinforced columns could
withstand large dynamic lateral loads and deflections. These RC columns had
shear strength higher than the shear corresponding to full development of
flexural strength under lateral load. Column longitudinal reinforcing bars had
sufficient anchorage and their strengths could be developed. The BPS team
observed that none of the columns failed in shear, and did not find any evidence
of pull-out of reinforcing bars. Longitudinal bars at ends of severely damaged
columns were observed to have fractured after necking (ductile failure). The BPS
team observed that severely bent columns did not seem to have been impacted
by a single, rigid object. Their deformed shapes were more consistent with loads
distributed over their heights. The observed deformation of columns, together
with the damage pattern throughout the building and the locations of aircraft
components suggest that the aircraft disintegrated rapidly as it entered the
building and collided with the first-floor columns. According to the BPS team, as
the moving debris from the aircraft pushed the contents and demolished exterior
wall of the building forward, the debris from the aircraft and building most likely
resembled a rapidly moving avalanche through the first floor of the building. The
BPS team concluded that the aircraft impact effects may be represented as a
rapid flow through the building of a fluid consisting of fuel and debris.
Two commercial airliners were crashed into the World Trade Center (WTC) North
and South Towers (WTC-I and WTC-II, respectively) in New York City, New
York, on September 11, 2001. Several studies on the response of WTC-I and
WTC-II to aircraft impact and/or thermal loading have been conducted in an
attempt to estimate the initial structural damage caused by the impact and
understand the collapse mechanism of the buildings under the ensuing fire
(FEMA, 2002; National Institute of Standards and Technology, NIST, 2005;

16
Baant and Zhou, 2002; Wierzbicki and Teng, 2003; Omika et al., 2005; Karim
and Fatt, 2005; Irfanoglu and Hoffmann, 2008). In the study conducted by NIST
(2005), computer simulations were used to investigate the effects of the impacts
of the airplanes on the buildings. Given the size and complexity of the structural
systems involved (WTC towers and aircrafts), simplifications had to be made to
reduce the size of the models used in global impact analyses. This was achieved
by conducting impact analyses of smaller element assemblies (referred to as
component impact analyses) to determine the sensitivity of impact response to
the modeling parameters. Component impact analyses showed that a normal
impact on the exterior wall of any of the WTC towers by an empty aircraft wing
section produced significant damage to the exterior columns (made of steel), but
not necessarily severance. For the same impact speed, component analyses
showed that impact by a fuel-filled aircraft wing section resulted in complete
failure (severance) of the exterior columns. NIST (2005) also found that the
debris resulting from impact of an aircraft wing section and an exterior-column
assembly, which propagated into the building, maintained most of the initial
momentum (before impact) of the impacting body.
The collision between a mass of liquid in a container and a structural barrier was
studied experimentally and analytically by Pujol and Brachmann (2007). This
problem was previously studied theoretically and numerically by Xue and
Wierzbicki (2003), who focused on high-speed impact of a liquid-filled cylinder
onto a rigid wall. The purpose of the study done by Pujol and Brachmann (2007)
was to determine a simple way to estimate the energy transferred to the barrier
by the impacting liquid body. This work focused on cases in which the barrier is
perpendicular to the trajectory of the liquid body, extends beyond the boundaries
of the area of intersection with the liquid, and retains its integrity during impact.
Eight tests were carried out using small, thin cylindrical containers filled with
beer. The containers were propelled through a barrel using compressed helium.
The barriers were square steel plates fixed at their four corners and instrumented

17
to measure the displacement at the center of the plate. In three out of eight tests,
the applied load was also measured. The velocity of the projectile before impact
was measured using two parallel laser beams and two photo-detectors monitored
with a data acquisition system. Projectile velocities up to approximately 200 mph
(90 m/s) were used. The data showed that the energy transferred to the barrier
can be estimated as the kinetic energy of the projectile before impact times the
ratio of the mass of the projectile, m, to the effective mass of the barrier, Me, if
m/Me << 1. This finding is relevant to the design of structural elements to resist
aircraft impact, as it has been shown that the fuel in the wings of an airplane can
cause severe damage to a structure (Mlakar et al., 2003; NIST, 2005).
Zineddin and Krauthammer (2007) tested RC slabs under impact load. Three
different series of tests were done to investigate the effects of steel reinforcement
and applied impact load on the dynamic response of the slabs. The support
conditions of the specimens were described as somewhere between simply
supported and fixed. A drop hammer device was used to apply the impact loads
at the centers of the slabs. The height from which the impacting mass was
dropped was changed from one test series to the next. The purpose of changing
the drop height was to investigate the influence of impact velocity on the failure
modes of the slabs (designed to fail in flexure under slowly applied load). The
specimens and the loading device were instrumented to obtain measurements of
impact loads, steel reinforcement strains, and slab deflections and accelerations.
High-speed videos were obtained to capture the events leading to failure of the
slabs. In several tests the slabs developed shear cracks and/or failed in punching
shear before significant bending and flexural cracks were observed. As the drop
height was increased, local response (in the impacted area) dominated the
behavior of the slab and the probability of punching or direct shear failure
increased. Slabs tested with the largest drop height failed in either punching
shear or direct shear. It was concluded that the favorable strength increase due
to the rate of loading can be offset by a change in the mode of failure.

18
2.4. Estimation of Maximum Dynamic Shear Demand in Beams
Biggs (1964) derived expressions for the maximum shear (support reaction) in a
beam loaded dynamically. Different beam support conditions, mass distributions,
load types, and response ranges were considered. The support reactions were
calculated considering the dynamic equilibrium of the beam. The shape of the
distribution of inertia forces along the span was assumed identical to the static
deflected shape of the beam. Beam resistance was defined in terms of the span
length and the internal moments developed. It was shown that the dynamic
reaction has two components: one that is a fraction of the total applied load, and
other that is a fraction of the resistance developed by the beam.
Keenan (1977) proposed a method to obtain the maximum shear in a one-way
slab subjected to blast load. The maximum dynamic shear was calculated as the
shear force associated with development of the ultimate flexural resistance of the
slab multiplied by a dynamic increase factor for maximum shear (DIFm). The
resistance function of the slab was assumed bilinear (elastic-perfectly plastic
behavior). A chart to evaluate DIFm was developed from the solutions of the
equations of motion of slabs with different stiffness and resistance, subjected to
triangular load pulses with different peak pressures and durations. The proposed
chart gives DIFm as a function of the ratios of load duration to fundamental period
of the slab and peak pressure to ultimate resistance of the slab. Following this
work, Murtha and Crawford (1981) developed a chart to obtain DIFm for simply
supported beams subjected to triangular pulses of uniformly distributed load.
The U.S. Army Corps of Engineers (1990) suggested that for short-duration blast
loads, the contribution of applied load to beam support shears can be neglected.
Thus, the maximum dynamic shear in a beam can be reasonably estimated as
the reaction force developed when the beam reaches its ultimate resistance. If
the ultimate resistance is not reached, the maximum dynamic shear is the
support shear corresponding to the maximum resistance attained by the element.

19
2.5. Estimation of Dynamic Shear Strength in RC Beams
The literature on experimental and analytical studies of the shear strength of RC
elements subjected to static load is rich (e.g., Richart, 1927; ACI-ASCE, 1962;
MacGregor, Sozen, and Siess, 1965; Bresler and MacGregor, 1967; Arakawa,
1969; Vecchio and Collins, 1986; Lynn, Moehle, Mahin, and Holmes, 1996; Pujol,
Sozen, and Ramirez, 2000; Hanai, Umemura, and Ichinose, 2005). Based on
experience and results from experimental and analytical research, ACI (2008)
proposed design criteria to prevent diagonal-tension and direct-shear (shearfriction concept) failure of RC beams under static load. Karagozian and Case
(1973) also proposed a design criterion to prevent direct-shear failure at the
support of an RC beam loaded statically. This criterion is less conservative than
the one proposed by ACI (2008).
Unlike the case of static loading, there are insufficient experimental data on the
dynamic shear strength of RC elements. Murtha and Crawford (1981) suggested
that the nominal static diagonal-tension and direct-shear strengths of RC beams
(ACI, 2008; Karagozian and Case, 1973) be increased by 50% to account for
dynamic effects on shear strength. The most relevant data on the shear strength
of RC elements subjected to blast load were reported by Slawson (1984). This
investigator proposed that in addition to the 50% dynamic increase suggested by
Murtha and Crawford (1981), increases of approximately 130% and 30% should
be applied to the nominal static diagonal-tension and direct-shear strengths,
respectively. Combining the dynamic strength increases mentioned above, it is
concluded from the work by Slawson (1984) that the diagonal-tension and directshear strengths of one-way RC slabs subjected to high-intensity blast load are
expected to be approximately 3 and 2 times, respectively, the corresponding
nominal static shear strengths.

20

CHAPTER 3. A HYPOTHESIS ON THE OCURRENCE OF INITIAL SHEAR


FAILURE IN BEAMS SUBJECTED TO BLAST LOAD

3.1. Introduction
The literature review presented in chapter 2 reveals consensus about a physical
phenomenon: a beam that fails in flexure under slowly applied load can fail in
shear under rapidly applied load. But there is no consensus about the cause of
this phenomenon. Some researchers have focused on the velocity imparted by
the load to the beam. Others have focused on force. All the approaches
reviewed, whether related to velocity or force, involve a large number of
parameters including parameters that are not usually expected to affect shear
failure such as the flexural strength of the beam.
The best evidence indicating a relationship between the mode of failure and the
loading rate of a beam comes from Menkes and Opat (1973). The results of their
tests are presented here in detail. To try to understand their observations, a
series of numerical simulations were done using the non-linear dynamic analysis
finite element (FE) program LS-DYNA (LSTC, 2005). The results from these
simulations are also described here. These results point towards a simple
hypothesis that is presented in the last section of this chapter.

21
3.2. Tests of Aluminum Beams Loaded with Explosives

3.2.1. General
Results from experiments on aluminum beams with fixed ends and subjected to
blast load have been reported by Menkes and Opat (1973). The beams had
different lengths and thicknesses. Similar beams showed different mechanisms
of failure depending on the intensity of the impulsive load. Beams subjected to
lower impulses showed major inelastic deformation. Beams subjected to higher
impulses failed in shear at the supports. Tearing (tensile failure) in outer fibers
at the ends of the beams occurred for intermediate impulses. This may be
interpreted to be the result of excessive rotation at plastic hinges near supports.
Properties of specimens, specific values of impulse, and test results are listed
next.

3.2.2. Experimental Plan


The experimental setup is shown in Figure 3.1. Aluminum beams with fixed ends
were subjected to high uniform pressure generated using sheet high explosive
(HE). A neoprene buffer was placed between the explosive and the top face of
the beam to distribute the impulse and prevent rear-surface spallation. The high
explosive (HE) used was Du Pont Detasheet D. To vary the load, sheets of HE
with four different thicknesses (10, 15, 25, and 50 mils) were combined. No
instrumentation was used. Only the permanent mid-span deflection was recorded
(when possible).
All beams were made of 6061-T6 aluminum. The cross section of the beams was
rectangular, with a width of one inch (b = 1). Beams with three different
thicknesses (h) were tested: 3/16, 1/4, and 3/8. Two different beam lengths
(clear span, L) were used: 4 and 8.

22
3.2.3. Experimental Results
Menkes and Opat (1973) did six series of tests (Table 3.1). Beams in a series
had the same geometry but were subjected to different explosive loads
generating different impulses. The nominal impulse per unit area (unit impulse, i)
imparted by the explosive load was computed using the empirical correlations
obtained by Clark (1965). In each series, the unit impulse i was increased from
test to test as shown in Table 3.1. Photographs taken after the tests of beams in
series 3 (Table 3.1) are shown in Figure 3.2. These photos show that as i was
increased, the failure mechanism (or failure mode) changed. Similar results were
observed for beams in the other series of tests. Three different failure modes
were identified:
I. Large inelastic deformation
II. Tearing (tensile failure) in outer fibers, at or over the support
III. Transverse shear failure at the support
The failure mode observed in each test is listed in Table 3.1. The extent of
damage to beams that showed Mode I is described by the magnitude of the
permanent deflection measured at mid-span, (Table 3.1). For any series, the
smallest unit impulse that caused tearing at the support was called Mode II
threshold. For a given series, the smallest unit impulse that caused shear failure
at the support was called Mode III threshold. The results indicate that for a
given beam thickness, the Mode II and III thresholds are independent of beam
length. These thresholds are summarized in Table 3.2 for the three different
beam thicknesses used. As the unit impulse i was increased above the Mode II
threshold, it was observed that Modes II and III overlapped and the beams failed
in a combination of tearing in the outer fibers and shear at the support. For
values of i above the Mode III threshold, the beams failed in a well-defined shear
mode, with the central portion of the span remaining flat.

23
3.3. Finite Element (FE) Simulation of Aluminum Beams Loaded Explosively

3.3.1. General
The general-purpose transient dynamic analysis FE program LS-DYNA (LSTC,
2005) was used to simulate the response of aluminum beams loaded with
explosives and tested by Menkes and Opat (1973) as described in section 3.2. A
comparison of the simulation results with the test data shows that LS-DYNA may
be an adequate tool to study the response of metal beams to blast load.
Simulation results for models of the beams in test series 3 (Table 3.1) are
presented next. Series 3 was selected for numerical simulation because it is the
series for which more detailed results are reported (Menkes and Opat, 1973).

3.3.2. Finite Element Model and Simulation Details


Eleven numerical models were developed to represent the aluminum beams
tested in series 3 (Table 3.1). The models were built using constant-stress solid
hexahedral elements. Single-node restraints were used to fix the ends of the
beam models. The material (6061-T6 aluminum) was represented using LS-DYNA
Material Type 24. This material model is included in LS-DYNA to represent an
isotropic elasto-plastic material with an arbitrary stress versus strain curve and
arbitrary strain rate dependency (LSTC, 2003). The true stress-true strain
relationship used in the models is shown in Figure 3.3a. This curve was
generated using experimental data (engineering stress/strain) obtained for 6061T651 aluminum (Figure 3.3b) by Davidson et al. (1975). The following material
properties were also used in the models:
Modulus of elasticity:

E = 10,000 ksi (69 GPa)

Yield stress (engineering):

Fy = 40 ksi (275 MPa)

Poissons ratio:

= 0.33

Mass density:

= 2.52 x 10-4 lbf-s2/in4 (2,700 kg/m3)

24
The failure criterion used was based on plastic strain. In this criterion, an element
is assumed to fail and is removed from the model when the plastic strain exceeds
a limit. This limit was fixed at 50% through calibration.
At strain rates of up to 10/s, no important increase in the tensile strength of 6061T6 aluminum has been observed (Hoge, 1966). Strain rates for the aluminum
beams described in section 3.2 are expected to be lower than 10/s and,
therefore, the material was regarded as strain-rate insensitive in the FE models.
Uniform pressure was applied to the models to represent the effect of the blast
load. The pressures that Menkes and Opat (1973) targeted were in the order of
kilobars (1 kbar ~ 14.5 ksi), and the load duration in the order of microseconds.
Triangular loads (Figure 3.4) with same duration td of 20 microseconds (0.02 ms)
and peak pressures p0 ranging from 15 to 90 ksi (1-6 kbar) as listed in Table 3.3
were assumed for analysis. The unit impulse (i = p0td/2) used in each FE
simulation matches the value listed in Table 3.1 for the corresponding test.
Calculated periods of vibration (T) of beams in all the series tested by Menkes
and Opat (1973) are listed in Table 3.4. For beams in series 3, T = 1.3 ms and
td/T = 0.02. The load applied to the beam models can be regarded as impulsive
(assumed to be the case for td < 0.1T).

3.3.3. Finite Element Simulation Results


Results from the LS-DYNA simulations of the response to blast load of the
beams tested in series 3 (Table 3.1) are presented in this section.
Renderings of final deformed shapes computed for beams in series 3 are shown
in Table 3.3. The variation of mid-span deflection with time computed for beams
in tests 1-3 (Mode-I deformation) is shown in Figure 3.5. Calculated permanent
deflections at mid-span () for beams in tests 1-3 are obtained from Figure 3.5

25
and listed in Table 3.3. The variation of maximum shear with time computed for
all beams in series 3 is shown in Figure 3.6a.

3.3.4. Discussion of Results


Final deformed shapes and/or failure mechanisms obtained experimentally and
numerically for the beams in series 3 are compared in Table 3.3. For a given test
with a certain unit impulse i, the beam deformation calculated using LS-DYNA is
similar to the deformation observed experimentally. Slight asymmetry in the
failure mode was observed in the actual tests for beams subjected to higher
impulses. This asymmetry is related to the fact that the explosive used to load
the beams was detonated from one end. This asymmetry was not observed in
the computed failure modes of beams in series 3 (Table 3.3). In the numerical
simulations pressure was assumed to be uniformly distributed.
The relative difference between computed and measured permanent deflection
at mid-span () ranged from 5% to 31%. Given that there is experimental error,
this range of variation seems acceptable. The experimental error is evidenced by
the fact that although the unit impulse was increased by 40% from test 2 to test
3, the permanent deflection reported for both tests was the same (Table 3.1).
Figure 3.6a shows that the maximum calculated shear force at the end of the
beam was reached within the time interval during which the load is acting on the
beam (0 < t < td). For the simulations of tests 1-4, as the unit impulse increased
(Table 3.3) the maximum shear force at the end of the beam increased as well.
For the simulations of tests 5-11, the maximum shear remained constant at
approximately 6.7 kip while the applied unit impulse increased by nearly 80%
(from 0.51 psi-s for test 5 to 0.89 psi-s for test 11). Beams in these tests were
reported to have failed near the supports and to have exhibited nearly zero
curvature at mid-span. Similar results for maximum calculated shear force were
obtained from simulations in which the duration of the load was 0.01 ms and the

26
peak pressures were twice the values indicated in Table 3.3 (the unit impulse in
each simulation, Table 3.3, remained constant), as shown in Figure 3.6b. The
observed limiting shear of 6.7 kip is close to the nominal plastic shear capacity of
the beams in Series 3:
Q0 = SA

Eq. 3.1

S = plastic shear strength


A = cross-sectional area
For the beams in series 3, A = bh = 0.25 in2 (Table 3.1). The plastic shear
strength S can be taken as 0 / 3 , where 0 is the strength of the material in
tension. For uniaxial loading, the maximum effective stress reached by the
material is equal to 0. From Figure 3.3b 0 = 50 ksi and, therefore, S = 29 ksi.
Using Eq. 3.1, the nominal plastic shear capacity (Q0) of the beams in Series 3
can be estimated as 7.2 kip. This is expected to be an upper bound to the actual
capacity of the beams in shear.
Considering that: (1) the computed limiting shear is close to the nominal plastic
shear capacity, (2) the nominal plastic shear capacity is an upper bound to shear
strength, and (3) the beams in which this limiting shear was reached did not
exhibit permanent bending near mid-span, it is reasonable to conclude that
beams 5-11 failed near the supports as the shear approached their plastic shear
capacities. This is confirmed by other analysis results that show that soon after
the beam reaches its capacity to resist shear, the reaction decreases to zero (as
elements that reach the effective plastic strain at failure are removed from the
model), deformations near the supports increase dramatically, and the middle of
the beam remains flat (Figure 3.7).

27
3.4. Hypothesis
Based on the experimental and numerical evidence presented in sections 3.2
and 3.3, respectively, a simple hypothesis is stated: A beam that fails in flexure
under static load may fail in shear under blast load if the maximum shear force
demand reaches the capacity of the beam to resist shear. Schematically:
Vmax Vn Shear failure

Eq. 3.2

Vmax = maximum shear demand in the initial phase of response to blast load
Vn = dynamic shear strength of the beam
For a beam that has been designed to develop a flexural mechanism under static
load, if the condition Vmax < Vn is satisfied for blast load, no change in the failure
mode (from flexure to shear) is expected and the designer may rely on high
energy-dissipation capacity through bending. The challenge in Eq. 3.2 is to
obtain appropriate estimates of the dynamic shear strength of the beam Vn, and
the maximum shear demand in the initial phase of response to blast load Vmax.
The effect of the loading rate (and therefore the strain rate) on the strength of the
beam material must be determined to calculate Vn. The shear demand Vmax
depends on the dynamics of the problem, in particular, on the initial deflected
shape of a beam subjected to blast load.
The hypothesis introduced above (Eq. 3.2) will be tested experimentally in
chapter 4 for beams with material properties (small-scale reinforced concrete)
and loading conditions (high-speed fluid impact) different from those used in the
tests described in section 3.2. This hypothesis will be also tested numerically in
chapter 6 for models of hypothetical beams subjected to rapidly applied uniform
pressure.

28
Table 3.1 Aluminum Beams Subjected to Blast Load Test Results
Series

Test

b
(in)

h
(in)

L
(in)

i
(psi-s)

Failure
Mode

(in)

1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
10
11
1
2
3
4
5
6
7
8
9
10
11

1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0

0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.187
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250
0.250

8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
4.0
4.0
4.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0

0.16
0.26
0.37
0.42
0.47
0.52
0.57
0.61
0.65
0.16
0.26
0.37
0.42
0.47
0.52
0.57
0.61
0.65
0.16
0.26
0.37
0.41
0.51
0.57
0.62
0.66
0.73
0.76
0.89
0.16
0.26
0.37
0.41
0.51
0.57
0.62
0.66
0.73
0.76
0.89

I
I
I
II
II
II-III
II-III
III
III
I
I
I
II
II
II-III
II-III
III
III
I
I
I
II
II-III
II-III
II-III
III
III
III
III
I
I
I
I
II-III
II-III
II-III
III
III
III
III

0.94
1.44
1.75
---*
---*
---*
---*
---*
---*
0.25
0.69
0.80
---*
---*
---*
---*
---*
---*
0.50
1.50
1.50
---*
---*
---*
---*
---*
---*
---*
---*
0.31
0.44
0.75
0.94
---*
---*
---*
---*
---*
---*
---*

* Beam was severed from supports

29
5

1
2
3
4
5
6
7
8
9
10
11
12
13
14
1
2
3
4
5
6
7
8
9
10
11
12
13
14

1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0

0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375
0.375

* Beam was severed from supports

8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
8.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0
4.0

0.26
0.31
0.37
0.41
0.48
0.51
0.57
0.62
0.73
0.82
0.89
0.92
0.95
1.01
0.26
0.31
0.37
0.41
0.48
0.51
0.57
0.62
0.73
0.82
0.89
0.92
0.95
1.01

I
I
I
I
I
I
I
I
II-III
II-III
II-III
II-III
III
III
I
I
I
I
I
I
I
I
II
II-III
II-III
III
III
III

Table 3.2 Mode II and Mode III Thresholds


h
(in)
0.187
0.250
0.375

Mode

i
(psi-s)

II
III
II
III
II
III

0.37
0.58
0.46
0.69
0.65
0.94

0.50
0.75
0.88
1.25
1.31
1.50
1.50
1.44
---*
---*
---*
---*
---*
---*
0.18
0.31
0.31
0.38
0.56
0.81
0.75
0.81
---*
---*
---*
---*
---*
---*

30
Table 3.3 Comparison of Experiments and Finite Element (FE) Simulations for
Beams in Series 3
Test

i,

psi-s

p0,
ksi

EXPERIMENT

Post-Test Photograph

LS-DYNA SIMULATION

,
in

At-Failure Rendering

,
in

0.16 15.7

0.50

0.59

0.26 25.6

1.50

1.03

0.37 36.8

1.50

1.58

0.41 41.3

0.51 50.5

0.57 57.0

0.62 61.7

0.66 66.2

0.73 73.0

10

0.76 76.1

11

0.89 88.6

Table 3.4 Natural Frequencies and Periods of Vibration of Aluminum Beams


Series
1
2
3
4
5
6

(rad/s)
3,730
14,910
4,980
19,930
7,470
29,890

T
(ms)
1.69
0.42
1.26
0.32
0.84
0.21

31

Figure 3.1 Experimental Setup - Tests on Aluminum Beams Loaded Explosively


(Menkes and Opat, 1973)

Figure 3.2 Results for Series of 6061-T6 Aluminum Beams (0.25x1x8-in. beam)
(after Menkes and Opat, 1973)

32

a) Effective Stress vs. Effective Plastic Strain LS-DYNA Input

b) Effective Stress vs. Effective Strain Experimental Data for 6061-T651


Aluminum (Davidson et. al. 1975)

Figure 3.3 Stress-Strain Relationship for Material Used in FE Models

Pressure

p0
Unit
impulse, i

td

Time

Figure 3.4 Load Pulse Used in FE Simulations

33

2.0

Midspan Deflection (in)

1.5

1.0

0.5
Test 1
Test 2
Test 3

0.0
0.0

0.2

0.4

0.6

0.8

1.0

Time (ms)

Figure 3.5 Calculated Midspan Deflection History Tests 1-3 in Series 3

34

Figure 3.6 Calculated Beam-End Shear Force History Series 3

35

t = 0.01 ms
Vend = 6.7 kip (max.)
t = 0.02 ms
Vend = 6.5 kip
t = 0.03 ms
Vend = 4.4 kip
t = 0.04 ms
Vend = 1.9 kip
t = 0.05 ms
Vend = 0 kip

Figure 3.7 LS-DYNA Results for the Variation with Time of End Shear (Vend) and
Deflected Shape as a Beam Subjected to Blast Load Fails near the Supports
(Test 8 in Series 3, load duration equal to 0.02 ms)

36

CHAPTER 4. EXPERIMENTS ON THE RESPONSE OF SMALL-SCALE


REINFORCED CONCRETE BEAMS TO FLUID IMPACT

4.1. Introduction
Tests were conducted to investigate whether the phenomenon observed by
Menkes and Opat (1973) can also occur for RC beams and for concentrated
dynamic loads. The tests were focused on the response of small-scale reinforced
concrete (RC) beams to impact of liquid masses traveling at speeds of up to 340
miles/hour. The setup shown in Figure A.21 was used. The impacting body
consisted of a thin-walled aluminum can filled with water. The test specimens
were small-scale concrete beams with longitudinal steel reinforcement and with
no transverse reinforcement.
A total of eleven series of tests were conducted. Eleven concrete batches were
made: one batch per test series. Properties measured for the concrete batches
made are listed in Table A.1. Geometry, material properties, and reinforcement
ratio of beams were identical in a given series, but varied from one series to the
next (Table 4.1). Series 1 had four beams, series 3, 5, 7-11 had five beams each,
and series 2, 4, 6 had one beam each, for a total of forty-two (42) test
specimens. One beam from each series was tested statically. The number of
specimens tested statically and dynamically in each series, and the numbers of
days that passed between casting and testing are listed in Table 4.2.
The beams had 1-in.-deep rectangular cross sections with widths ranging from 2
to 6 in. (Figure A.14). The beams had end rotations restrained (Figure 4.1) and
were loaded at midspan. Two different clear spans were used in the tests: 8 and
12 in. Assuming full rotational restraint at the ends, the shear spans (a) were 3 in.

37
and 2 in. for beams in series 1-10 and 11, respectively (Table 4.1). The effective
depth (d) of all the beams was 0.75 in., resulting in a/d ratios of 4 and 2.7 in
series 1-10 and 11, respectively. The compressive strength of concrete (at age of
testing) was in the range 5-10 ksi. The longitudinal reinforcement consisted of
continuous top and bottom layers of 0.1-in. diameter steel wire (Figure A.14). The
reinforcement was heat treated using two different procedures that led to two
average values of yield stress: 120 ksi and 50 ksi. Two different reinforcement
ratios were used: 0.5% and 1% (top and bottom). Appendix A gives details of the
dimensions and material properties of the beams, and the fabrication procedure.

4.2. Static Tests

4.2.1. General
The goal of the experimental program described in this chapter was to test the
hypothesis formulated in section 3.4 using small-scale RC beams subjected to
concentrated high-speed fluid impact load. The hypothesis states that a beam
that fails in flexure under static load may fail in shear under rapidly applied load if
the initial shear demand reaches the capacity of the element to resist shear.
Therefore, the mechanisms of failure of the test specimens had to be established
for the case of static load before impact tests were conducted. One beam from
each series was tested statically using the setup shown in Figure A.20, for a total
of eleven static tests. The beams were subjected to monotonically increasing
concentrated load at mid-span (Figure 4.2). Typical pretest and posttest views of
the beams in the static tests in series 1, 3, 5, 7-11 are shown in Figure 4.3.
Details of the static test setup, the support conditions of the beams, and the
instrumentation and testing procedure used are given in Appendix A.

38
4.2.2. Load-Deflection Response
The applied load and beam deflections were measured in each test. The load
was controlled up to cracking. After cracking, beam deflection was controlled.
The goal of these tests was simply to determine the failure mechanism under
static load. For series 1-7, the tests were carried out until reductions in lateral
resistance (applied load) of 20-40% of peak loads were observed. To investigate
the possibility of catenary action, the tests in series 8-11 were carried out until
the beams reached drift ratios of approximately 20-25% (the drift ratio, , is
defined in Figure 4.2) and/or the loading mechanism became unstable.
Beam deflections were measured at the quarter points and at midspan using
linear variable differential transformers (LVDTs, see Appendix A) as shown in
Figure A.22. Deflection at midspan was also measured using a dial gage. Loadmidspan deflection curves of the beams tested are shown in Figure 4.4 to Figure
4.14. Deflected shapes at peak loads (listed in Table 4.3) are shown in Figure
4.15 to Figure 4.20. Cracking loads (listed in Table 4.3) were estimated after the
tests using the measured load-deflection curves of the beams. Cracks in all
beams formed only at the sections of maximum positive and negative moment
(i.e., midspan and ends of the beam, as shown in Figure 4.3), and were initially
difficult to observe. As the deflections of the beams increased, cracks widened.
Initial beam stiffness (obtained from the measured load-deflection curve), drift
ratio at peak load ((Pmax)), and maximum drift ratio (max) in each test are listed in
Table 4.3. The full flexural capacity of each beam was reached in each test,
except for the beams in series 2, 4, 6, which failed in shear.

4.2.3. Failure Mechanisms


Details of the specimens after failure are shown in Figure 4.21 to Figure 4.31.
From these figures and from the load-deflection curves presented in Figure 4.4 to
Figure 4.14 it is clear that under static load, the beams in series 1, 3, 5, 7-11
developed flexural mechanisms of failure and beams in series 2, 4, 6 failed in

39
shear. Beams that developed flexural mechanisms of failure reached high drift
ratios (max > 7%, Table 4.3; the drift ratio was defined in Figure 4.2). Their
deformations concentrated at sections near the supports and at midspan. These
deformations were related to one wide crack (exceeding 1/16 in width) at each
critical section. This crack was nearly perpendicular to the longitudinal axis of the
beam. Beams that failed in shear developed wide inclined cracks (exceeding
1/16 in width) near midspan at low drift ratios ((Pmax) < 3%) and had narrow
flexural cracks (not exceeding 0.02 in. in width).

4.2.4. Support Rotations


The supports were designed to restrain the ends of the beams against rotation
(Figure 4.1). Details of the dimensions of the support plates and the clamping
force applied at each support are given in Appendix A. The upper support of the
specimen was instrumented to measure rotation in series 7-11 (Appendix A).
Estimates of maximum support rotations ((support)max) in the static tests in series
7-11 are listed in Table 4.4. This table also lists the ratios of maximum support
rotation to maximum drift ratio. The maximum ratio was less than 0.002 (0.2%).

4.3. Impact Tests

4.3.1. General
In the static tests (section 4.2), beams from eight series (1, 3, 5, 7-11) developed
flexural mechanisms of failure and beams from three series (2, 4, 6) failed in
shear. Beams from the eight series in which the beams developed flexural
mechanisms under static load were tested dynamically (impact test). Three
impact tests were done in series 1 and four impact tests were done in each of the
other seven series (Table 4.5), for a total of thirty-one (31) impact tests. The
setup shown in Figure A.21 was used in these tests. For simplicity, specimens

40
are designated using two numerals that refer to the experimental series (Table
4.1) to which the specimen belongs, and the impact test ID within a particular
series (Table 4.5). For instance, the specimen from series 1 used in the test
identified in Table 4.5 as test 2 of this series is referred to as specimen S1_T2.
The beams were impacted at mid-span by cylindrical aluminum containers filled
with water (Figure 4.32) and traveling perpendicular to the beams (Figure 4.33).
Containers with the same dimensions (Appendix A) were used in all tests. The
mass of water in the container was approximately 0.4 lb (181 g) in all tests. The
total mass of the projectile (container and water) was approximately 0.42 lb
(190 g). The impact velocity was changed from one test to the next in each series
to impart different forces to the beams. The projectiles were accelerated to
speeds of up to 340 miles/hour (150 m/s). A high-speed video camera (section
A.4.5) was used to record the collision of the projectile and the specimen. Typical
pretest and posttest views of the beams tested under impact are shown in Figure
4.34. Details of the test setup, beam support conditions, instrumentation, and
testing procedure used are given in Appendix A.

4.3.2. Measured Impact Velocities


Projectile velocity (or impact velocity), vp, was measured using two different
methods. One method involved the use laser beams and photodetectors. The
other involved the use of the high-speed camera. Both methods are described in
Appendix A. Impact velocities measured in all tests using these methods are
listed and compared in Table 4.5. The maximum deviation in measured vp
between the two methods was 4%. Velocities obtained using the laser beams
and photodetectors were deemed more reliable (Appendix A). The results and
discussions presented in chapter 5 refer to these velocities.

41
4.3.3. Response to Impact and Failure Mechanisms
The extent and type of damage caused by impact changed from test to test,
depending on beam geometry, reinforcement ratio, reinforcement yield stress,
and impact velocity. Details of all the specimens after impact are shown in Figure
4.35 to Figure 4.65. All beams experienced cracking and inelastic deformations.
Cracking patterns were reconstructed from images captured with the high-speed
camera described in Appendix A. The cracking patterns of all the beams are
shown in Figure 4.66. Maximum and permanent midspan deflections (max and
perm, respectively) measured in the impact tests are listed in Table 4.6.
Maximum and permanent drift ratios (max and perm, respectively) are also listed
in Table 4.6 (drift ratio, , is defined in Figure 4.33).
The state of the specimens after impact ranged from complete disintegration to
small permanent deflections accompanied by narrow cracks (not exceeding 0.03
in. in width). Thirteen specimens had the same deformation/failure modes under
impact and static load (flexural mechanisms). Eighteen specimens had failure
modes under impact load that were different from the failure modes under static
load. In these specimens failure under impact load was dominated by inclined
cracking. Damage observed in the impact tests is summarized in Table 4.6.
Series 1
Posttest photographs of the test specimens in series 1 are shown in Figure 4.35
to Figure 4.37.
Specimen S1_T1 (Figure 4.35) had narrow flexural cracks (not exceeding 0.03
in. in width) and small deflections. Three hairline cracks perpendicular to the
beam axis formed: one near each support and one at midspan. Maximum and
permanent midspan deflections were 0.29 in. and 0.17 in., respectively.

42
Specimen S1_T2 (Figure 4.36) showed inclined cracking which led to failure. In
the upper half of the beam, a hairline inclined crack formed from midspan to
approximately the upper quarter point. In the lower half of the beam, a wide
inclined crack extended from midspan to approximately the lower quarter point. A
crack perpendicular to the beam axis formed at midspan (Figure 4.66). A wedge
delimited by the crack at midspan and the inclined crack in the lower half of the
beam was blown off the beam. Buckling of the reinforcement was observed at
the location of this wedge. The beam was severed from the lower support. A
crack perpendicular to the beam axis and fracture of one reinforcing wire were
observed near the upper support.
Specimen S1_T3 (Figure 4.37) was partially disintegrated (losing approximately
40% of its mass) during impact. The remainder of the beam was severed from
the lower support. High-speed video revealed that concrete disintegration started
with the formation of an inclined crack in the upper third of the specimen (Figure
4.66). Photographic frames extracted from the high-speed videos recorded
during the tests and showing the sequences of events leading to beam
disintegration are presented in section 4.3.4. Cracks perpendicular to the beam
axis formed at midspan and near the upper support. A wide crack (exceeding
0.08 in. in width) across the full beam depth formed near the lower quarter point.
Series 3
Posttest photographs of the test specimens in series 3 are shown in Figure 4.38
to Figure 4.41.
Specimen S3_T1 (Figure 4.38) developed a flexural mechanism of failure.
Maximum and permanent midspan deflections were 1.6 in. and 1.5 in.,
respectively. Wide cracks (exceeding 3/16 in width) and plastic hinges formed
near the supports and at midspan. The cracks near the supports were flexureshear cracks.

43
Specimen S3_T2 (Figure 4.39) had inclined cracking and partial disintegration
(losing approximately one third of its mass during impact), but seemed to have
developed a flexural mechanism of failure. An inclined crack formed near
midspan. Wide flexure-shear cracks (exceeding 1/4 in width) formed near the
supports. Concrete at midspan spalled off the beam. Fracture of reinforcement
was observed near the upper support and at midspan.
Specimen S3_T3 (Figure 4.40) developed a flexural mechanism of failure.
Maximum midspan deflection could not be measured because LVDTs detached
from the beam. Permanent midspan deflection was approximately 2.3 in. Wide
cracks (exceeding 1/4 in width) and plastic hinges formed near the supports
and at midspan. A hairline crack parallel to the beam axis formed on the front
face of the specimen from midspan to the upper support (Figure 4.40b-c).
Fracture of one reinforcing wire near the upper support and spalling of concrete
in the area of impact were observed.
Specimen S3_T4 (Figure 4.41) had inclined cracking and partial disintegration
(losing approximately 60% of its mass) during impact. High-speed video revealed
that concrete disintegration started with the formation of two inclined cracks.
These cracks extended from midspan to the upper quarter point and from
approximately the lower quarter point to the lower support (Figure 4.66). Wide
flexure-shear cracks (exceeding 1/8 in width) formed near the supports.
Series 5
Posttest photographs of the test specimens in series 5 are shown in Figure 4.42
to Figure 4.45.
Specimen S5_T1 (Figure 4.42) experienced nearly total disintegration (losing
approximately 80% of its mass) during impact. The remainder of the beam was
severed from the lower support. An inclined crack formed near the lower quarter

44
point. High-speed video revealed that concrete disintegration started with
punching failure (section 4.3.4) in the center portion of the beam. A wide
flexure-shear crack formed near the upper support (Figure 4.66).
Specimen S5_T2 (Figure 4.43) experienced nearly total disintegration (losing
approximately 70% of its mass) during impact. The remainder of the beam was
partially severed from the upper support. High-speed video showed that concrete
disintegration started with punching failure (section 4.3.4) in the center portion
of the beam. Tow inclined cracks formed: one near the upper quarter point and
the other near the lower support (Figure 4.66). A crack parallel to the beam axis
formed on the front face of the specimen along one of the reinforcing wires
(Figure 4.43b-c). This crack suggests bond failure.
Specimen S5_T3 (Figure 4.44) had narrow flexural cracks (not exceeding 0.02
in. in width) and small deflections. A hairline crack perpendicular to the beam
axis formed near each support. Maximum and permanent midspan deflections
were 0.24 in. and 0.07 in., respectively.
Specimen S5_T4 (Figure 4.45) had narrow flexural cracks (not exceeding 0.02
in. in width) and small deflections. A hairline crack perpendicular to the beam
axis formed near each support. Maximum and permanent midspan deflections
were 0.22 in. and 0.08 in., respectively.
Series 7
Posttest photographs of the test specimens in series 7 are shown in Figure 4.46
to Figure 4.49.
Specimen S7_T1 (Figure 4.46) experienced wide flexural cracks (exceeding 1/8
in width) and large deflections. Maximum and permanent midspan deflections
were 1.2 in. and 1.1 in., respectively. The cracks formed near the supports and at

45
midspan, and were nearly perpendicular to the beam axis. A plastic hinge
formed at midspan.
Specimen S7_T2 (Figure 4.47) developed a flexural mechanism of failure.
Plastic hinges formed near the supports and at midspan. A wide flexure-shear
crack (exceeding 1/4 in width) formed near each support. Fracture of reinforcing
wires was observed near the supports and at midspan. The reinforcing wire
closest to the front face of the beam was pulled out of the concrete near
midspan. Concrete on the back face of the specimen spalled off at midspan.
Specimen S7_T3 (Figure 4.48) was partially disintegrated (losing approximately
40% of its mass) during impact. High-speed video revealed that concrete
disintegration started with the formation of an inclined crack in the lower third of
the beam (Figure 4.66). This crack had a length of approximately 3 in. (or L/4,
where L is the clear span). Two plastic hinges and two wide cracks (with widths
exceeding 1/4) nearly perpendicular to the beam axis formed near the supports.
Fracture of reinforcing wires was observed near the supports and at midspan.
Specimen S7_T4 (Figure 4.49) developed a flexural mechanism of failure.
Maximum and permanent midspan deflections were 1.5 in. and 1.4 in.,
respectively. Plastic hinges and wide cracks (exceeding 3/16 in width) formed
near the supports and at midspan. The cracks near the supports were flexureshear cracks.
Series 8
Posttest photographs of the test specimens in series 8 are shown in Figure 4.50
to Figure 4.53.
Specimen S8_T1 (Figure 4.50) had narrow flexural cracks (not exceeding 0.03
in. in width) and small deflections. Three hairline cracks nearly perpendicular to

46
the beam axis formed: one near each support and one at midspan. Maximum
and permanent midspan deflections were 0.29 in. and 0.19 in., respectively.
Spalling of concrete was observed in the area of impact.
Specimen S8_T2 (Figure 4.51) had inclined cracking and bond failure. Highspeed video showed that bond failure started with the formation of two inclined
cracks in the upper half of the beam. These cracks extended from midspan to the
upper quarter point (Figure 4.66). A longitudinal crack formed along the bottom
reinforcement and extended from the upper quarter point to the upper support.
Maximum and permanent midspan deflections were 1.0 in. and 0.95 in.,
respectively. Wide flexure-shear cracks (exceeding 1/16 in width) formed near
the supports. Spalling of concrete was observed in the area of impact.
Specimen S8_T3 (Figure 4.52) experienced inclined cracking and bond failure.
High-speed video revealed that bond failure started with the formation of two
inclined cracks in the upper half of the beam. The inclined cracks formed near
the upper support and midspan (Figure 4.66). A longitudinal crack formed along
the bottom reinforcement. Permanent midspan deflection was 0.90 in. A wide
crack (exceeding 1/16 in width) nearly perpendicular to the beam axis formed
near each support. Spalling of concrete was observed in the area of impact.
Specimen S8_T4 (Figure 4.53) was partially disintegrated (losing approximately
30% of its mass) during impact. High-speed video revealed that concrete
disintegration started with the formation of an inclined crack in the upper third of
the specimen (Figure 4.66). An inclined crack formed near midspan in the lower
half of the beam. Flexure-shear cracks formed near the supports (Figure 4.66).
Series 9
Posttest photographs of the test specimens in series 9 are shown in Figure 4.54
to Figure 4.57.

47
Specimen S9_T1 (Figure 4.54) experienced wide flexural cracks (exceeding 1/8
in width) and large deflections. One crack formed near each support and one
crack formed at midspan. Maximum and permanent midspan deflections were
1.2 in. and 1.1 in., respectively.
Specimen S9_T2 (Figure 4.55) was disintegrated during impact. High-speed
video revealed that concrete disintegration started with the formation of an
inclined crack and a longitudinal crack in the lower third of the beam (Figure
4.66). A flexure-shear crack formed near the upper support and an inclined crack
formed near the lower support (Figure 4.66).
Specimen S9_T3 (Figure 4.56) developed a flexural mechanism of failure.
Maximum midspan deflection could not be measured because LVDTs detached
from the beam. Permanent midspan deflection was 1.5 in. Wide cracks
(exceeding 3/16 in width) and plastic hinges formed near the supports and at
midspan. Spalling of concrete was observed in the area of impact.
Specimen S9_T4 (Figure 4.57) was disintegrated during impact. High-speed
video revealed that concrete disintegration started with the formation of multiple
inclined cracks near the supports, the quarter points, and midspan (Figure 4.66).
Bond failure seems to have taken place in the upper third of the beam.
Series 10
Posttest photographs of the test specimens in series 10 are shown in Figure 4.58
to Figure 4.61.
Specimen S10_T1 (Figure 4.58) had flexural cracks and small deflections. One
crack formed near each support. Two slightly inclined hairline cracks formed near
midspan: one in the upper half and one in the lower half of the beam. Maximum
and permanent midspan deflections were 0.32 in. and 0.25 in., respectively.

48
Specimen S10_T2 (Figure 4.59) had wide inclined cracks (exceeding 1/16 in
width) and large deflections. Two inclined cracks formed in the lower half of the
beam: one extending from midspan to the lower quarter point, and one near the
support. An inclined crack formed in the upper half of the beam and extended
from midspan to the upper quarter point. Maximum midspan deflection could not
be measured because LVDTs detached from the beam. Permanent midspan
deflection was 0.90 in. A flexure-shear crack formed near the upper support and
a flexural crack formed at midspan. A hairline crack parallel to the beam axis
formed on the front face of the specimen from midspan to the upper quarter point
(Figure 4.59d). Spalling of concrete was observed in the area of impact.
Specimen S10_T3 (Figure 4.60) had inclined cracks and large deflections. Two
inclined cracks formed in the upper and lower halves of the beam. These cracks
extended from midspan to approximately the quarter points. Maximum midspan
deflection could not be measured because LVDTs detached from the beam.
Permanent midspan deflection was 0.80 in. Flexural cracks formed near the
supports and at midspan. Cracks parallel to the beam axis formed on the front
face of the specimen (Figure 4.60b). These cracks extended across the full beam
length. Spalling of concrete was observed in the area of impact.
Specimen S10_T4 (Figure 4.61) had inclined cracks and partial disintegration
(losing approximately 50% of its mass) during impact. High-speed video revealed
that concrete disintegration started with the formation of inclined cracks near the
quarter points and midspan (Figure 4.66). A wide flexural crack (exceeding 1/8
in width) formed near the upper support. A flexure-shear crack formed near the
lower support. A crack parallel to the beam axis formed on the front face of the
specimen along one of the reinforcing wires (Figure 4.61b). This crack suggests
bond failure.

49
Series 11
Posttest photographs of the test specimens in series 11 are shown in Figure 4.62
to Figure 4.65.
Specimen S11_T1 (Figure 4.62) had flexural cracks (approximately 1/16 wide)
and small deflections. One crack formed near each support and one crack
formed at midspan. They were nearly perpendicular to the beam axis. Maximum
and permanent midspan deflections were 0.39 in. and 0.30 in., respectively.
Specimen S11_T2 (Figure 4.63) had inclined cracks and partial disintegration
(losing approximately 60% of its mass) during impact. High-speed video revealed
that concrete disintegration started with the formation of inclined cracks near the
supports and the quarter points (Figure 4.66). The beam was severed from the
upper support. Fracture of reinforcement was observed near the upper support.
Specimen S11_T3 (Figure 4.64) experienced inclined cracking and partial
disintegration (losing between 60-70% of its mass) during impact. High-speed
video revealed that concrete disintegration started with the formation of two
inclined cracks (Figure 4.66). One inclined crack formed in the upper half of the
beam. The other inclined crack formed in the lower third of the beam and was
approximately 3-in. long (or 3L/8). The beam was partially severed from the
upper support. Fracture of reinforcement was observed near the upper support.
Specimen S11_T4 (Figure 4.65) was disintegrated during impact. High-speed
video revealed that concrete disintegration started with the formation of an
inclined crack in the upper third of the beam and a longitudinal crack in the lower
half of the beam (Figure 4.66). The longitudinal crack formed along the top
reinforcement (suggesting bond failure). High-speed video showed that flexureshear cracks formed near the supports (Figure 4.66). Fracture of reinforcement
was observed near the upper support.

50
4.3.4. Initial Damage in Impact Tests
Damage observed in the impact tests was described in section 4.3.3. It was
mentioned that in some tests high-speed video revealed the formation of inclined
cracks that led to failure/disintegration of the specimens (Table 4.6). These
cracks formed at early stages of response and were typically located between
the quarter points and midspan. It was also mentioned in section 4.3.3 that highspeed video revealed initial punching failures leading to disintegration of two
specimens (S5_T1 and S5_T2). Series of consecutive photographic frames
taken during impact were extracted from high-speed videos recorded in different
tests. This was done to identify the punching failure as well as the formation and
propagation of the inclined cracks mentioned before, and to understand the
mechanisms leading to failure and/or disintegration of the specimens.
The series of photographic frames obtained are shown in Figure 4.67 to Figure
4.74. A given series shows the sequence of events in a given impact test at early
times (typically within 2 ms after impact). A time stamp is provided below each
photographic frame. All frames show side views of the specimens as the position
of the camera remained constant during the test. The specimen is approximately
centered in the vertical direction in each frame. The position of the specimen is
indicated in Figure 4.67. The projectile travels from left to right in the photos and
is identified in Figure 4.67. Cardboard or foamboard placed aside the specimen
to keep water away and increase the clearness of high-speed imagery (section
A.4.4) is also shown in the photos. The cardboard or foamboard appears
between the projectile and the specimen in the photos and is identified in the
figures. In Figure 4.70 to Figure 4.74 the upper end of the specimen is not visible.
Relevant aspects of initial response to impact (inclined cracks, punching or bond
failure) captured in the high-speed videos are circled in the photo frames for
easier observation.

51
4.3.5. Deflection Histories and Deflected Shapes
LVDTs (Appendix A) were installed behind the beams to measure deflections at
the quarter points and midspan (Figure A.22). The cores of the LVDTs were
attached to the back face of the beams using aluminum brackets (Figure A.23c).
Deflection measurements could not be obtained in tests in which the beams
disintegrated (Table 4.6). Similarly, deflection measurements could not be
obtained in tests in which the concrete on the back face of the beams was lost
through spalling or disintegration. The variations of midspan deflection with time
for specimens that did not disintegrate or spalled (Table 4.6) are presented in
Figure 4.75 to Figure 4.82. The permanent midspan deflection observed in each
deflection history matches the midspan deflection measured directly after the
test. Deflected shapes at the times of maximum midspan deflection are shown in
Figure 4.83 to Figure 4.88.

4.3.6. Support Rotations


The supports were designed to restrain rotations at the ends of the beams
(Figure 4.1). Details of the dimensions of the support plates and the clamping
force applied at each support are presented in Appendix A. The upper support of
the specimen was instrumented to measure support rotation in series 7-10
(Appendix A). It was difficult to obtain credible estimates of support rotations in
tests in which the specimens disintegrated because of impact (Appendix A).
Estimates of maximum support rotations ((support)max) in the impact tests in series
7-10 are listed in Table 4.7. This table also lists the ratios of maximum support
rotation to maximum drift ratio. The maximum ratio was approximately 0.02 (2%).

52

Table 4.1 Properties of Beams in the Different Experimental Series


Steel Reinforcement
Beam Dimensions
Compressive
Strength of Nominal Yield
Clear Effective
Reinforcement
Concrete,
Series
Strength,
Depth,
Width,
b
Depth,
h
Span,
Ratio,
=
(MM/DD/YYYY)
fc(1)
(top and bottom)
fy
d=d
(in)
(in)
L
(psi)
(%)
(ksi)
(in)
(in)
Date of
Casting

(1)

Shear
Span,
a/d
a
(in)

01/08/2009

9200

120

0.5

12

0.75

01/07/2009

8500*

120

1.0

12

0.75

01/09/2009

7630

120

0.5

12

0.75

01/11/2009

8410

120

1.0

12

0.75

01/12/2009

9120

120

0.5

12

0.75

01/15/2009

9860

120

1.0

12

0.75

06/19/2009

7040

50

0.5

12

0.75

06/22/2009

6330

50

1.0

12

0.75

06/24/2009

6390

50

0.5

12

0.75

10

06/25/2009

7690

50

1.0

12

0.75

11

06/29/2009

7720

50

0.5

0.75

2.7

Measured compressive strength of concrete at age of static (series 2, 4, 6) or impact (series 1, 3, 5, 7-11) tests (Table A.1 and Table A.2).

* Compressive strength (fc) was not measured in series 2 because cylinders were defective (section A.2.1). The average of fc at age of static tests in series

52

1, 3-6 was used.

53

Table 4.2 Experimental Program Number of Specimens Tested


Series

Static Tests
Number of Days
Number of
between Casting
Beams Tested
and Testing

Impact Tests
Number of Days
Number of
between Casting
Beams Tested
and Testing

103

123-130

96

---

---

96

133-153

95

---

---

97

108-118

95

---

---

45

79-86

43

85-91

43

92-98

10

43

99-107

11

43

56-66

Table 4.3 Summary of Results for Beams Tested Statically

Series

Cracking
Initial
Stiffness, Load,
Pcr
k0
(lbf)
(kip/in)

Peak
Load,
Pmax
(lbf)

100

280

740

86

280

1060

3
4
5

160
170
145

540
540
500

180

750

75

200

50

200

100

350

10
11

115
300

500
780

1270
2140
1740
2960
350
580*
820
1280
1370

Deflection
at Peak
Load,
(Pmax)
(in)
0.24
0.15
0.22
0.18
0.34
0.16
0.21
0.16
0.18
0.16
0.13

Drift Ratio
Maximum Maximum
at Peak
Deflection, Drift Ratio,
Load,
max
max
(Pmax)
(in)
(rad)
(rad)
0.039

0.42

0.070

0.025

0.16

0.027

0.037

0.62

0.103

0.029

0.28

0.047

0.056

0.45

0.075

0.027

0.20

0.033

0.035

0.92

0.153

0.026

1.25

0.208

0.030

1.50

0.250

0.026

1.44

0.240

0.032

0.97

0.243

* First peak load. When the midspan deflection exceeds 0.5 in., the resistance of the beam starts
increasing up to a maximum load of approximately 670 lbf (catenary action).

54
Table 4.4 Maximum Support Rotations in Static Tests Series 7-11
Maximum Support Maximum Drift
Rotation, (support)max
Ratio, max
(support)max/max
(rad)
(rad)
5.6 x 10-5
0.153
0.0004
-4
2.4 x 10
0.208
0.0012
-4
2.8 x 10
0.250
0.0011
1.5 x 10-4
0.240
0.0006
-4
3.7 x 10
0.243
0.0015

Series
7
8
9
10
11

Table 4.5 Measured Impact Velocities


Series
1
3

Impact (Projectile) Velocity, vp


(vp)camera/
Test ID Lasers/photodetectors High-speed camera
(vp)lasers
(in/s)
(m/s)
(in/s)
(m/s)
1
2
3
1
2
3
4

1
2
3
4

1
2
3
4

1
2
3
4

1
2
3
4

10

1
2
3
4

11

1
2
3
4

1860
3370
3670
3430
4150
3950
4390
5560
5770
3260
2950
2210
2590
3230
2930
1840
2630
2280
3050
2570
3260
3150
4150
2780
3740
3460
4210
3020
3780
4550
5100

47.2
85.6
93.2
87.1
105.4
100.3
111.5
141.2
146.6
82.8
74.9
56.1
65.8
82.0
74.4
46.7
66.8
57.9
77.5
65.3
82.8
80.0
105.4
70.6
95.0
87.9
106.9
76.7
96.0
115.6
129.5

1800
3400
3600
3460
4190
4000
4380
5620
5620
3160
2900
2220
2570
3210
2900
1840
2610
2250
3000
2570
3210
3100
4090
2730
3600
3460
4090
3100
3750
4500
5290

45.7
86.4
91.4
87.9
106.4
101.6
111.3
142.7
142.7
80.3
73.7
56.4
65.3
81.5
73.7
46.7
66.3
57.2
76.2
65.3
81.5
78.7
103.9
69.3
91.4
87.9
103.9
78.7
95.3
114.3
134.4

0.97
1.01
0.98
1.01
1.01
1.01
1.00
1.01
0.97
0.97
0.98
1.00
0.99
0.99
0.99
1.00
0.99
0.99
0.98
1.00
0.98
0.98
0.99
0.98
0.96
1.00
0.97
1.03
0.99
0.99
1.04

55

Table 4.6 Midspan Deflections and Damage in Impact Tests


Specimen
S1_T1
S1_T2
S1_T3
S3_T1
S3_T2
S3_T3
S3_T4

Impact
Velocity,
vp
(in/s)

Maximum
Deflection,
max
(in)

1860
3370
3670
3430
4150
3950
4390
5560
S5_T1
5770
S5_T2
S5_T3
3260
S5_T4
2950
2210
S7_T1
2590
S7_T2
S7_T3
3230
S7_T4
2930
1840
S8_T1
2630
S8_T2
S8_T3
2280
S8_T4
3050
2570
S9_T1
3260
S9_T2
S9_T3
3150
S9_T4
4150
2780
S10_T1
3740
S10_T2
S10_T3
3460
S10_T4
4210
3020
S11_T1
3780
S11_T2
S11_T3
4550
S11_T4
5100
* Beam disintegrated.

0.29
---*
---*
1.6
---*
---**
---*
---*
---*
0.24
0.22
1.2
---***
---*
1.5
0.29
1.0
---**
---*
1.2
---*
---**
---*
0.32
---**
---**
---*
0.39
---*
---*
---*

Maximum Permanent
Drift Ratio, Deflection,
max
perm
(rad)
(in)
0.048
---*
---*
0.267
---*
---**
---*
---*
---*
0.040
0.037
0.200
---***
---*
0.250
0.048
0.167
---**
---*
0.200
---*
---**
---*
0.053
---**
---**
---*
0.098
---*
---*
---*

0.17
---*
---*
1.5
---*
2.3
---*
---*
---*
0.07
0.08
1.1
---***
---*
1.4
0.19
0.95
0.90
---*
1.1
---*
1.5
---*
0.25
0.90
0.80
---*
0.30
---*
---*
---*

** Concrete on the back face of the beam spalled off.


*** Concrete at midspan was blown off the beam.

Permanent
Drift Ratio,
perm
(rad)
0.028
---*
---*
0.250
---*
0.383
---*
---*
---*
0.012
0.013
0.183
---***
---*
0.233
0.032
0.158
0.150
---*
0.183
---*
0.250
---*
0.042
0.150
0.133
---*
0.075
---*
---*
---*

Observed Damage
Disintegration
of Beam
No
Yes
Yes
No
Yes
No
Yes
Yes
Yes
No
No
No
No
Yes
No
No
No
No
Yes
No
Yes
No
Yes
No
No
No
Yes
No
Yes
Yes
Yes

Inclined
Cracking
No
Yes
Yes
No
Yes
No
Yes
Yes
Yes
No
No
No
No
Yes
No
No
Yes
Yes
Yes
No
Yes
No
Yes
No
Yes
Yes
Yes
No
Yes
Yes
Yes

56
Table 4.7 Maximum Support Rotations in Impact Tests Series 7-10
Maximum Support Maximum
Rotation,
Drift Ratio,
(support)max/max
Specimen
(support)max
max
(rad)
(rad)
-3
0.200
S7_T1
2.7 x 10
0.014
S7_T2
---*
---*
---*
S7_T3
---*
---*
---*
0.250
S7_T4
3.6 x 10-3
0.014
0.048
S8_T1
6.0 x 10-4
0.012
-3
0.167
S8_T2
0.015
2.5 x 10
S8_T3
2.8 x 10-3
0.150**
0.019**
S8_T4
4.1 x 10-3
---***
---***
-3
0.200
S9_T1
3.5 x 10
0.018
S9_T2
---*
---*
---*
-3
S9_T3
4.7 x 10
0.250**
0.019**
S9_T4
---*
---*
---*
-3
0.053
S10_T1
1.1 x 10
0.021
-3
0.150**
S10_T2
1.9 x 10
0.013**
0.133**
S10_T3
2.0 x 10-3
0.015**
S10_T4
---*
---*
---*
* Support rotation could not be estimated. LVDT 4 went out of range because it was
impacted by water and concrete debris.
** Permanent drift ratio was used instead of maximum drift ratio (not available because
maximum midspan deflection was not measured).
*** Deflection measurements were not obtained (specimen disintegrated) and drift ratio
could not be computed.

57

Figure 4.1 Beam Supports

Deflected shape

L/2

= /(L/2)

Figure 4.2 Schematic Description of Static Test and Drift-Ratio Definition

58

Figure 4.3 Typical Views of Beams in Static Tests in Series 1, 3, 5, 7-11

59

Figure 4.4 Applied Load vs. Beam Deflection at Midspan Static Test_Series 1

Figure 4.5 Applied Load vs. Beam Deflection at Midspan Static Test_Series 2

60

Figure 4.6 Applied Load vs. Beam Deflection at Midspan Static Test_Series 3

Figure 4.7 Applied Load vs. Beam Deflection at Midspan Static Test_Series 4

61

Figure 4.8 Applied Load vs. Beam Deflection at Midspan Static Test_Series 5

Figure 4.9 Applied Load vs. Beam Deflection at Midspan Static Test_Series 6

62

Figure 4.10 Applied Load vs. Beam Deflection at Midspan Static Test_Series 7

Figure 4.11 Applied Load vs. Beam Deflection at Midspan Static Test_Series 8

63

Figure 4.12 Applied Load vs. Beam Deflection at Midspan Static Test_Series 9

Figure 4.13 Applied Load vs. Beam Deflection at Midspan Static Test_Series 10

64

Figure 4.14 Applied Load vs. Beam Deflection at Midspan Static Test_Series 11

65

Figure 4.15 Deflected Shape of Beam at Peak Load Static Tests_Series 1-2

Figure 4.16 Deflected Shape of Beam at Peak Load Static Tests_Series 3-4

Figure 4.17 Deflected Shape of Beam at Peak Load Static Tests_Series 5-6

66

Figure 4.18 Deflected Shape of Beam at Peak Load Static Tests_Series 7-8

Figure 4.19 Deflected Shape of Beam at Peak Load Static Tests_Series 9-10

Figure 4.20 Deflected Shape of Beam at Peak Load Static Test_Series 11

67

Figure 4.21 Damage and Condition of Beam at End of Static Test Series 1

Figure 4.22 Damage and Condition of Beam at End of Static Test Series 2

68

Figure 4.23 Damage and Condition of Beam at End of Static Test Series 3

Figure 4.24 Damage and Condition of Beam at End of Static Test Series 4

69

Figure 4.25 Damage and Condition of Beam at End of Static Test Series 5

Figure 4.26 Damage and Condition of Beam at End of Static Test Series 6

70

Figure 4.27 Damage and Condition of Beam at End of Static Test Series 7

Figure 4.28 Damage and Condition of Beam at End of Static Test Series 8

71

Figure 4.29 Damage and Condition of Beam at End of Static Test Series 9

Figure 4.30 Damage and Condition of Beam at End of Static Test Series 10

72

Figure 4.31 Damage and Condition of Beam at End of Static Test Series 11

73

Figure 4.32 Container Used in Impact Tests

Deflected shape

vp

L
Projectile

L/2

= /(L/2)

Figure 4.33 Schematic Description of Impact Test

74

Figure 4.34 Typical Views of Beams in Impact Tests

75

Figure 4.35 Damage and Condition of Beam after Impact Specimen S1_T1

Figure 4.36 Damage and Condition of Beam after Impact Specimen S1_T2

76

Figure 4.37 Damage and Condition of Beam after Impact Specimen S1_T3

77

Figure 4.38 Damage and Condition of Beam after Impact Specimen S3_T1

Figure 4.39 Damage and Condition of Beam after Impact Specimen S3_T2

78

Figure 4.40 Damage and Condition of Beam after Impact Specimen S3_T3

Figure 4.41 Damage and Condition of Beam after Impact Specimen S3_T4

79

Figure 4.42 Damage and Condition of Beam after Impact Specimen S5_T1

Figure 4.43 Damage and Condition of Beam after Impact Specimen S5_T2

80

Figure 4.44 Damage and Condition of Beam after Impact Specimen S5_T3

Figure 4.45 Damage and Condition of Beam after Impact Specimen S5_T4

81

Figure 4.46 Damage and Condition of Beam after Impact Specimen S7_T1

Figure 4.47 Damage and Condition of Beam after Impact Specimen S7_T2

82

Figure 4.48 Damage and Condition of Beam after Impact Specimen S7_T3

Figure 4.49 Damage and Condition of Beam after Impact Specimen S7_T4

83

Figure 4.50 Damage and Condition of Beam after Impact Specimen S8_T1

Figure 4.51 Damage and Condition of Beam after Impact Specimen S8_T2

84

Figure 4.52 Damage and Condition of Beam after Impact Specimen S8_T3

Figure 4.53 Damage and Condition of Beam after Impact Specimen S8_T4

85

Figure 4.54 Damage and Condition of Beam after Impact Specimen S9_T1

Figure 4.55 Damage and Condition of Beam after Impact Specimen S9_T2

86

Figure 4.56 Damage and Condition of Beam after Impact Specimen S9_T3

Figure 4.57 Damage and Condition of Beam after Impact Specimen S9_T4

87

Figure 4.58 Damage and Condition of Beam after Impact Specimen S10_T1

Figure 4.59 Damage and Condition of Beam after Impact Specimen S10_T2

88

Figure 4.60 Damage and Condition of Beam after Impact Specimen S10_T3

Figure 4.61 Damage and Condition of Beam after Impact Specimen S10_T4

89

Figure 4.62 Damage and Condition of Beam after Impact Specimen S11_T1

Figure 4.63 Damage and Condition of Beam after Impact Specimen S11_T2

90

Figure 4.64 Damage and Condition of Beam after Impact Specimen S11_T3

Figure 4.65 Damage and Condition of Beam after Impact Specimen S11_T4

91

Specimen

Typical

S1-T1

S1-T2

S1-T3

S3-T1

S3-T2

S3-T3

S3-T4

S5-T1

S5-T2

S5-T3

S5-T4

S7-T1

S7-T2

S7-T3

S7-T4

S8-T1

S8-T2

S8-T3

S8-T4

S9-T1

S9-T2

S9-T3

S9-T4

S10-T1

S10-T2

S10-T3

S10-T4

S11-T1

S11-T2

S11-T3

S11-T4

Cracking
Pattern

Specimen

Cracking
Pattern

Specimen

Cracking
Pattern

Figure 4.66 Cracking Patterns of Beams in Impact Tests

92

Figure 4.67 Initial Damage in Impact Tests Specimen S1_T2

92

93

Figure 4.68 Initial Damage in Impact Tests Specimen S1_T3

93

94

Figure 4.69 Initial Damage in Impact Tests Specimen S5_T2

94

95

Figure 4.70 Initial Damage in Impact Tests Specimen S8_T2

95

96

Figure 4.71 Initial Damage in Impact Tests Specimen S9_T4

96

97

Figure 4.72 Initial Damage in Impact Tests Specimen S10_T3

97

98

Figure 4.73 Initial Damage in Impact Tests Specimen S10_T4

98

99

Figure 4.74 Initial Damage in Impact Tests Specimen S11_T3


99

100

1.8
1.6
1.4

Deflection (in)

1.2
1.0
0.8
0.6
0.4
0.2
0.0
0.000
-0.2

Time (s)
0.005

0.010

0.015

0.020

Figure 4.75 Midspan Deflection History Specimen S1_T1

1.8
1.6
1.4

Deflection (in)

1.2
1.0
0.8
0.6
0.4
0.2
0.0
0.000
-0.2

Time (s)
0.005

0.010

0.015

Figure 4.76 Midspan Deflection History Specimen S3_T1

0.020

101

1.8
1.6
1.4

Deflection (in)

1.2
1.0
0.8
S5_T3 (vp = 3260 in/s)

0.6

S5_T4 (vp = 2950 in/s)

0.4
0.2
0.0
0.000
-0.2

0.005

Time (s)

0.010

0.020

0.015

Figure 4.77 Midspan Deflection History Specimens S5_T3, S5_T4

1.8
1.6
1.4

Deflection (in)

1.2
1.0
S7_T1 (vp = 2210 in/s)

0.8

S7_T4 (vp = 2930 in/s)

0.6
0.4
0.2
0.0
0.000
-0.2

Time (s)
0.005

0.010

0.015

Figure 4.78 Midspan Deflection History Specimens S7_T1, S7_T4

0.020

102

1.8
1.6
1.4

Deflection (in)

1.2
1.0
0.8
S8_T1 (vp = 1840 in/s)

0.6

S8_T2 (vp = 2630 in/s)

0.4
0.2
0.0
0.000
-0.2

Time (s)
0.005

0.010

0.015

0.020

Figure 4.79 Midspan Deflection History Specimens S8_T1, S8_T2

1.8
1.6
1.4

Deflection (in)

1.2
1.0
0.8
0.6
0.4
0.2
0.0
0.000
-0.2

Time (s)
0.005

0.010

0.015

Figure 4.80 Midspan Deflection History Specimen S9_T1

0.020

103

1.8
1.6
1.4

Deflection (in)

1.2
1.0
0.8
0.6
0.4
0.2
0.0
0.000
-0.2

Time (s)
0.005

0.010

0.015

0.020

Figure 4.81 Midspan Deflection History Specimen S10_T1

1.8
1.6
1.4

Deflection (in)

1.2
1.0
0.8
0.6
0.4
0.2
0.0
0.000
-0.2

Time (s)
0.005

0.010

0.015

Figure 4.82 Midspan Deflection History Specimen S11_T1

0.020

104

Distance from Support (in)


6

12

0.0

Deflection (in)

0.3
0.6
0.9
1.2
1.5
1.8

Figure 4.83 Deflected Shape at Time of Maximum Deflection Specimen S3_T1

Distance from Support (in)


6

12

0.0

Deflection (in)

0.3
0.6
0.9
S7_T1 (vp = 2210 in/s)

1.2

S7_T4 (vp = 2930 in/s)

1.5
1.8

Figure 4.84 Deflected Shape at Time of Maximum Deflection Specimens


S7_T1, S7_T4

Distance from Support (in)


6

12

0.0

Deflection (in)

0.3
0.6
0.9
1.2
1.5
1.8

Figure 4.85 Deflected Shape at Time of Maximum Deflection Specimen S8_T1

105

Distance from Support (in)


6

12

0.0

Deflection (in)

0.3
0.6
0.9
1.2
1.5
1.8

Figure 4.86 Deflected Shape at Time of Maximum Deflection Specimen S9_T1

Distance from Support (in)


6

12

0.0

Deflection (in)

0.3
0.6
0.9
1.2
1.5
1.8

Figure 4.87 Deflected Shape at Time of Maximum Deflection Specimen S10_T1

Distance from Support (in)


4

0.0

Deflection (in)

0.3
0.6
0.9
1.2
1.5
1.8

Figure 4.88 Deflected Shape at Time of Maximum Deflection Specimen S11_T1

106

CHAPTER 5. DISCUSSION OF EXPERIMENTAL RESULTS

5.1. Introduction
The experimental program conducted was described in chapter 4. The
specimens in the impact tests (section 4.3) showed different failure modes. Some
specimens had large deflections and developed flexural mechanisms of failure.
Others had inclined cracking and disintegrated during impact (Table 4.6). In a
given series, as the impact velocity (and, therefore, the impact force exerted to
the beam) increased the failure mode of the beams changed although beams in
a series had the same nominal properties. Flexural mechanisms of failure were
observed at low impact velocities. Inclined cracking leading to disintegration of
the specimens was observed at high impact velocities. Control specimens
developed flexural mechanisms of failure under static load (section 4.2). To
understand the reason for a change in the failure mode, the observed response
of the specimens is studied in this chapter.

5.2. Static Tests

5.2.1. General
The results of the static tests were presented in section 4.2. Calculated values of
initial stiffnesses, cracking loads, and peak loads are presented and compared
with the test results in this section. A discussion on the maximum shear resisted
by the beams and their nominal shear strengths under static load is presented in
this section as well.

107
5.2.2. Initial Stiffness and Cracking Load
Values of initial stiffness (before cracking), k0, and cracking load, Pcr, were
obtained from the measured load-deflection responses (Figure 4.4 to Figure
4.14) and are listed in Table 4.3. Assuming full rotational restraint at the ends of
the beam and assuming the applied load to be concentrated at midspan, k0 and
Pcr are given by:
k0 =

192 E c I g

Pcr =

Eq. 5.1

8M cr
L

Eq. 5.2

Ec = modulus of elasticity of concrete


Ig = gross moment of inertia (uncracked section)
L = clear span
Mcr = cracking moment
Eq. 5.2 is obtained by considering the moment diagram of a fixed-fixed beam
with a concentrated load at midspan (Figure 5.1). The cracking moment Mcr is:

M cr =

fr I g

Eq. 5.3

fr = modulus of rupture of concrete


Ig = gross moment of inertia (uncracked section)
c = distance from neutral axis to extreme fiber in tension at the section of
maximum moment (c = h/2, where h is the depth of the beam)
All beams had rectangular cross sections with h = 1 in. and, therefore, c = 0.5 in.
The gross moment of inertia of the beam sections is Ig = bh3/12. Values of Ig and
L are listed in Table 5.1 for beams in all series.

108
To calculate the initial stiffness (Eq. 5.1) and the cracking moment (Eq. 5.3) of a
beam in the static tests (section 4.2), the moduli of elasticity (Ec) and rupture (fr)
of concrete are required. The modulus of elasticity was measured for concrete
batches 7-11 (section A.2.1) after all impact tests in series 7-11 were completed.
The ages of the cylinders at the days when Ec was measured are listed in Table
A.3. The ages of the beams at the days of static tests are listed in Table 4.2.
For a given series, the date of the static test and the date when Ec was measured
differed by as much as 82 days (Table A.3 and Table 4.2).
Mean measured values of Ec and fc (compressive strength) for concrete batches
7-11 are listed in Table A.1. Figure A.10 in Appendix A shows that the variation
of Ec with fc for the small-scale concrete mix used is represented reasonably well
(maximum deviation from mean measured values is less than 20%) with the
expression suggested by ACI-318 (2008) for regular concrete:

Ec = 57,000 f 'c

Eq. 5.4

where Ec and fc are in psi. The modulus of elasticity of the small-scale concrete
was computed using this expression and the estimated compressive strength at
the day of the static test. This strength was estimated for each series from curves
showing the variation of fc with time (Figure A.1 to Figure A.4). Estimates of fc at
the days of static tests are listed in Table 5.1. Using these values of fc and Eq.
5.4, estimates of Ec at the days of static tests were obtained (Table 5.1).
Barry and Sozen (1970) studied the variation of strength with time for a smallscale concrete mix similar to the one used in the experiments described here. It
was observed that the ratio of the modulus of rupture to the compressive
strength increased with time. At 12 days after casting, the modulus of rupture (fr)
was observed to be approximately 10 f ' c [psi]. At 48 days after casting, fr was

109

observed to be approximately 15 f 'c [psi]. Given the similarity of the mixes, the
test results obtained by Barry and Sozen were used to estimate fr at the days of
static tests. Static tests were done between 95-103 days after casting in series 16, and between 43-45 days after casting in series 7-11 (Table 4.2). The modulus
of rupture is expected to be close to 15 f 'c [psi] in series 7-11. The time interval
over which fr needs to be estimated in series 1-6 is significantly different from the
time interval studied by Barry and Sozen. There is no data for fr beyond 48 days
after casting. However, fr is not expected to increase indefinitely and 15 f 'c [psi]
may be adopted as an upper limit. This value would apply to unreinforced beams
(specimens typically tested to obtain fr). The beams in the static tests have
longitudinal reinforcement and, therefore, their moduli of rupture are expected to
be lower than 15 f 'c [psi] because of restrained shrinkage. It is believed that

12 f 'c [psi] may be a more realistic estimate of fr for the beams in the static
tests. This relationship was used to calculate the values of fr listed in Table 5.1.
Using the estimated values of fr at the days of static tests, the cracking moments
(Mcr) of the beams were calculated using Eq. 5.3 and are listed in Table 5.1. The
cracking loads (Pcr) were obtained from the estimated values of Mcr (Eq. 5.3)
using Eq. 5.2 and are also listed in Table 5.1. Values of initial beam stiffness
were calculated using Eq. 5.1 and the estimated values of Ec at the days of static
tests, and are listed in Table 5.1.
Ratios of calculated cracking load to measured cracking load and calculated
initial stiffness to measured initial stiffness are presented in Table 5.1 for all
series. The mean ratios are approximately 1.0 and 1.4 for cracking load and
initial stiffness, respectively. The error associated to calculation of initial beam
stiffness is not unacceptable in view of the uncertainties involved. These
uncertainties are related to the estimation of Ec at the day of static test and to the
assumption of full rotational restraint at the supports. There is also uncertainty in

110
determining the cracking loads from the measured load-deflection responses as
initial cracking was difficult to observe (section 4.2.2). Shear deformations have
been assumed to be negligible and the calculated initial stiffnesses were based
on flexural deformations only. This may lead to calculated stiffnesses higher than
measured stiffnesses.

5.2.3. Peak Load


Peak loads (Pmax) measured in the static tests are listed in Table 4.3. Assuming
full rotational restraint at the ends of the beam and assuming the applied load to
be concentrated at midspan, Pmax can be expressed as:

Pmax =

8M max
L

Eq. 5.5

L = clear span
Mmax = maximum bending moment that the beam section can develop
Eq. 5.5 can be obtained by considering the moment diagram of a fixed-fixed
beam subjected to concentrated load at midspan (Figure 5.1). The maximum (or
limiting) moment Mmax in each series was calculated using the spreadsheet
FLECHA (Pujol, 2001). The stress-strain relationship (SSR) of the concrete was
represented with Hognestads parabola. Estimates of the compressive
sytrength of the concrete (fc) at the days of static tests (Table 5.1) were used in
the calculations. A limiting strain in the concrete of 0.004 was also used in these
calculations (concrete had no confinement because of the absence of transverse
reinforcement in the test specimens). This strain limit comes from experience
related to regular concrete. Its applicability here, for small-scale concrete, is
admittedly debatable. The limit is simply used here for lack of a better estimate.
The SSR of the reinforcement (Figure A.12-Figure A.13) was represented with an

111
elasto-plastic relationship. Strain hardening was considered for beams in series
7-11 (with SSR as shown in Figure A.13). No axial load was input in FLECHA.
Moment-curvature relationships obtained with FLECHA for the beam sections in
all the test series are shown in Figure 5.2 to Figure 5.6. The peak moment values
from these curves (Mmax) are listed in Table 5.2. The peak loads (Pmax) were
obtained from the values of Mmax using Eq. 5.5 and are also listed in Table 5.2.
Ratios of calculated peak loads to measured peak loads are listed in Table 5.2
for all series. The mean ratio is 0.77. The differences between measured and
calculated strengths (Pmax) can be attributed at least in part to friction in the
supports. The impossibility to eliminate friction between the steel rollers and the
steel plates while limiting rotation at the upper support of the specimen (Figure
4.1) caused axial restraint of the beam. When subjected to transverse load, the
beam initially elongates. If the axial (vertical) deformation of the specimen is
partially restrained, axial force is induced. This axial force increases the flexural
capacity of the section. The induced axial force can be estimated as two times
the total clamping force at each support (24 kips, section A.4.1) multiplied by a
friction coefficient . The multiplier of 2 is used because there were two friction
surfaces at the upper support (Figure 4.1). For = 0.025 (2.5%), the induced
axial force is 1.2 kip. This force is enough to cause increases in the flexural
capacities of the beams sufficiently large to explain the differences between
measured and calculated peak loads. Calculated values of Mmax (from FLECHA)
and Pmax (Eq. 5.5) accounting for possible axial forces on the specimens (with
= 0.025) are listed in Table 5.2. The mean ratio of calculated Pmax including the
effect of axial restraint to measured Pmax is 1.05.

5.2.4. Maximum Shear


The maximum shear force in a test beam is Vmax = Pmax/2 (Figure 5.1), where
Pmax is the peak load measured in the static test (Table 4.3). Maximum shear

112
forces in the beams in the static tests are listed in Table 5.2. The maximum
nominal (average) shear stress (max) in a beam is:

max =

Vmax
bd

Eq. 5.6

b = beam width
d = effective depth of beam
Vmax = maximum shear force in the beam
Maximum nominal shear stresses (max) in the beams tested statically are listed
in Table 5.2. In series 1, 3, 5, 7-11, max was lower than 3 f 'c (the highest max
was 2.7 f 'c in series 11), where max and fc are in psi units. Beams in these
series developed flexural mechanisms of failure under static load (section 4.2). In
series 2, 4, 6, max was higher than 3 f 'c . Beams in these series failed in shear
under static load (section 4.2). The mean static shear strength of the beams in
series 2, 4, and 6 was 3.7 f 'c [psi]. The nominal static shear strength, Vn [lbf],
of the beams in series 2, 4, and 6 can be computed as:

Vn = 3.7 f 'c bd

Eq. 5.7

fc = compressive strength of concrete (psi)


b = beam width (in)
d = effective depth of beam (in)
Chana (1981) conducted a series of static tests on small-scale reinforced
concrete (RC) beams similar to the test specimens in this study. The beams
tested by Chana had simple supports and rectangular cross sections. The
sections were 0.9-in. wide and 1.8-in. deep (the effective depth was 1.6 in.). The

113
clear span was 14.4 in. Two point loads were used. Shear-span-to-depth ratio
was 3.0. Longitudinal reinforcement ratio was 1.8%. No transverse reinforcement
was provided. Compressive strength of concrete ranged from 5.2 ksi to 9.3 ksi.
Yield stress of steel reinforcement was approximately 80 ksi. All beams failed in
shear at nominal shear stresses n ranging from 3.3 f ' c to 4.6 f ' c [psi].
The results by Chana (1981) and the results from series 2, 4, 6 suggest that it is
reasonable to assume the static shear strength of beams in series 1, 3, 5, 7-11 to
be close to 3.7 f ' c [psi].

5.3. Impact Tests

5.3.1. General
The results of the impact tests were presented in section 4.3. Idealizations of the
load and specimens used in these tests will be introduced in this section.
Calculated periods of vibration of beams, magnitude and duration of load, and
maximum deflections of specimens that experienced flexural deformation in the
impact tests are presented. Calculated beam deflections are compared with
deflections measured in the impact tests. A discussion on the maximum initial
shear experienced by the beams and their dynamic shear strengths is presented
in this section as well.

5.3.2. Dynamic Idealization of Specimens using SDOF Systems


The study of the response of a structural element to dynamic load can be
simplified by idealizing both the load and the element (Biggs, 1964). A beam
subjected to impact can be represented with an equivalent single-degree-offreedom (SDOF) system (Figure 5.7). The SDOF system has a spring constant ke

114
(stiffness), a mass Me, and applied load Fe. These parameters are selected so
that the deflection of the equivalent SDOF system is equal to the deflection of the
beam at midspan (where the test specimen was impacted). Following the work
by Biggs (1964), the relationships between the mass, stiffness, and applied load
in the SDOF system and their counterparts in the beam can be written as:
M e = KM M

Eq. 5.8

ke = K L k

Eq. 5.9

Fe = K L F

Eq. 5.10

Me = mass of equivalent SDOF system


ke = spring constant (stiffness) of equivalent SDOF system
Fe = load applied to equivalent SDOF system
M = beam mass
k = beam stiffness
F = load applied to beam
KM = mass factor
KL = load (stiffness) factor
The factors KM and KL depend on the beam support conditions, mass distribution,
load type, and type of response (linear or non-linear). A deformed shape for the
beam must be assumed to derive the mass and load factors. The deformed
shape under static load is typically used for this purpose. The factor KM is
obtained by equating the kinetic energy (KE) of the equivalent SDOF system to
the KE of the beam. The factor KL is obtained by equating the strain energy (SE)
of the equivalent SDOF system with the SE of the beam. The stiffness factor is
identical to KL and is obtained by equating the work done by the equivalent
applied load Fe on the SDOF system with the work done by the applied load F on
the beam. Biggs (1964) provided detailed information about the mass and load
factors for equivalent SDOF systems to represent beams and slabs. For a fixed-

115
fixed beam with uniform mass, concentrated load at midspan, and responding in
the linear range the mass and load factors are:

K M = 0.37

Eq. 5.11

K L = 1.0

Eq. 5.12

For a fixed-fixed beam with uniform mass, concentrated load at midspan, and
responding in the plastic range the mass and load factors are:

K M = 0.33

Eq. 5.13

K L = 1.0

Eq. 5.14

Assuming different deformed shapes, the mass and load factors can be obtained
for beams with different types of loads and boundary conditions (Biggs, 1964).
Using Eq. 5.11 and Eq. 5.13, the masses of the equivalent SDOF systems for
linear and plastic response of the specimens in the impact tests are obtained
(Table 5.3). From Eq. 5.12 and Eq. 5.14, the stiffnesses of the equivalent SDOF
systems are equal to the stiffnesses of the beams in the impact tests. The
resistance of each SDOF system used to represent a test specimen was
idealized as shown in Figure 5.8, where:
ke = effective stiffness (spring constant),
Rm = maximum resistance, and
y = yield deflection.
The parameters ke and Rm were selected to represent the load-deflection
responses of the beams measured in the static tests (Figure 4.4 to Figure 4.14)
and are listed in Table 5.3. The parameter y is calculated as:

116

y =

Rm
ke

Eq. 5.15

Rm = maximum resistance
y = yield deflection
Values of y for the equivalent SDOF systems used (Table 5.3) were obtained
using Eq. 5.15 and the values of ke and Rm listed in Table 5.3.
The period of vibration of the equivalent SDOF systems is equal to the effective
period of the beam and is given by:

T = 2

Me
ke

Eq. 5.16

Me = mass of equivalent SDOF system


ke = spring constant (stiffness) of equivalent SDOF system
The fundamental period of vibration of each test specimen (Table 5.3) was
computed using Eq. 5.16 and the values of Me (in the linear range of response)
and ke listed in Table 5.3. If the initial stiffness k0 is used in Eq. 5.16 instead of ke,
the calculated uncracked periods of vibration (Tuncr.) of the beams are obtained.

5.3.3. Idealization of Impact Load


The interaction between a deformable barrier and an impacting liquid body is a
complex phenomenon. This problem has been previously studied experimentally
and analytically by other researchers (Sugano et al, 1993; Xue and Wierzbicki,
2003; Pujol and Brachmann, 2007). These studies dealt with the impact of a
liquid-filled body on a flat barrier. The force exerted on the flat surface (F) can be
described using the expression proposed by Riera (1968):

117
F (t ) = Fcr [ x(t )] + [ x(t )] v(t ) 2

Eq. 5.17

t = time
x(t) = position along the length of the impacting body (perpendicular to
the flat surface)
Fcr = load needed to crush or buckle the body containing the liquid,
expressed as a function of x

= mass per unit length of the impacting body


v = relative velocity between the impacting body and the barrier
In cases in which the container is light and weak, Fcr can be neglected. This is
done here to estimate the force demand on the beams in the impact tests
described in section 4.3. This assumption is justified by the small mass of the
container (approximately 9 g, which represents less than 5% of the total mass of
the projectile; section 4.3.1) and its low load-carrying capacity (the thickness of
the container shell is 0.002 in; section A.4.3).
Because the strength of the container has been neglected (Fcr = 0), the force
exerted on the beams depends exclusively on and v. The second term on the
right-hand side of Eq. 5.17 is obtained from conservation of linear momentum in
the direction in which the impacting body travels (Figure 4.33). For analysis
purposes, the mass of the container will be also neglected and the impacting
body is taken as a cylinder of water with mass m (181 g, section 4.3.1), crosssectional area A, and length lw. Thus, the mass per unit length of the impacting
body is constant throughout its length and can be expressed as:

[ x(t )] = = w A

Eq. 5.18

w = density of water (1000 kg/m3 = 9.35 x 10-5 lbf-s2/in4)


A = cross-sectional area of impacting cylinder of water (3.14 in2; the inner
diameter of the container is 2 in.)

118
The length of the cylinder of water is lw = m/[wA] = 3.5 in. (the total length of the
container is 3.75 in). If the velocity of the impacting body is high enough to
produce a load that acts on the beam for a small fraction of its period (Table 5.3),
it can be assumed that no beam deformation occurs within the duration of the
load. In this case, the relative velocity between the impacting body and the beam
(v) is equal to the projectile velocity, vp (section 4.3.2, Table 4.5). The impact load
is constant with time and can be idealized as shown in Figure 5.9. The duration td
and magnitude F0 (using Eq. 5.17-5.18) of the idealized impact load are:
td = lw / v p

Eq. 5.19

Eq. 5.20

F0 = w Av p

lw = length of impacting cylinder of water


vp = projectile velocity (impact velocity)

w = density of water
A = cross-sectional area of impacting cylinder of water
If the duration of the load is short, it is assumed that the beam retains its integrity
during impact, and that initially there are no deformations and strain energy. The
energy transferred from the liquid to the beam is initially kinetic energy only. This
initial kinetic energy of the beam (KE0) can be estimated using the expression
proposed by Pujol and Brachmann (2007):

KE 0 =

m
KE p
Me

Eq. 5.21

m = mass of impacting liquid body (181 g = 0.4 lb = 1.03 x 10-3 lbf-s2/in)


Me = effective beam mass (or mass of equivalent SDOF system)
KEp = kinetic energy of impacting liquid body

119
Eq. 5.21 was proposed for impact durations and masses of impacting liquid
bodies within the ranges td T/4 and m 0.06Me, respectively. In Eq. 5.21, the
kinetic energy of the impacting liquid body (KEp) is:

KE p =

1
2
mv p
2

Eq. 5.22

vp = projectile velocity (impact velocity)


Using Eq. 5.19-5.22, the duration (td) and magnitude (F0) of the idealized load,
the kinetic energy of the projectile (KEp), and the kinetic energy transferred to the
beam (KE0) in each impact test (Table 4.6) are obtained. Values of td, F0, KEp,
and KE0 are listed in Table 5.4. The ratios m/Me (using Me in the linear range)
required to calculate KE0 (Eq. 5.22) are listed in Table 5.3. For all specimens in
the impact tests, m/Me > 0.17. The mass ratios m/Me in the impact tests are out
of the range studied by Pujol and Brachmann (2007).
Ratios of duration of the idealized load to calculated effective period of vibration
of the beam (td/T) in each impact test are listed in Table 5.4. The values of td/T
range approximately from 0.4 to 1.4 and are out of the range studied by Pujol
and Brachmann (2007). The durations of the impact loads are long compared to
the periods of the beams and the idealization of the load shown in Figure 5.9
might not be appropriate.
Cracking and spalling of concrete are expected to occur as the specimens are
impacted by the liquid bodies. As a result, the periods of vibration (T) of the
beams will lengthen and, therefore, the ratios td/T will decrease. The lengthening
of the beam periods could be as high as 100% of the values listed in Table 5.4,
producing values of td/T ranging between 0.2 and 0.7. Even in this case, td is
relatively long and the applicability of Eq. 5.19-5.21 to study the overall dynamic

120
response of the beams is questionable. The loads in the impact tests are better
described as dynamic rather than impulsive. Commonly, a load is assumed to be
impulsive if td < 0.1T (Biggs, 1964).
For m and td approaching Me and T, respectively, the interaction between the
liquid and the beam can be studied in an incremental fashion. The force exerted
on the beam by a small portion of liquid (acting for a short time on the beam) can
be estimated using Eq. 5.20. In the impact tests, several beams (18 out of 31)
had inclined cracking (section 4.3, Table 4.6). The inclined cracks formed early in
the response of the beams to the impact loads (section 4.3.4). It is believed that
the force calculated using Eq. 5.20 can be used to estimate the initial shear
stresses causing the inclined cracks in the beams.

5.3.4. Maximum Initial Shear


To estimate initial shear demand on the test specimens, consider the impact of a
small segment of the liquid body on the beam (Figure 5.10a). The impacting
segment of liquid has cross-sectional area A, length x, mass m, density w,
and velocity vp (in a direction perpendicular to the beam). Because x << lw, the
loading time associated with the segment of liquid with mass m is short
compared to the period of the beam. Therefore, Eq. 5.20 can be used to estimate
the maximum force initially exerted on the beam by the impacting liquid (F0). This
force is assumed to be applied in the middle of the beam over a circular area
equal to the cross-sectional area of the impacting liquid body (A), which has a
diameter of 2 in. The critical beam sections for shear demand are expected to be
located at 1 to 2 in. from midspan (sections AA and BB in Figure 5.10b). The
dynamic equilibrium of the beam portion between sections AA and BB (ABBA)
is considered to estimate shears at the critical sections. The free body diagram of
ABBA is shown in Figure 5.10c. Because of the small mass of ABBA, the initial
inertia force related to this portion of the beam is assumed to be negligible.
Therefore, translational equilibrium requires:

121
F0 = V AA' + VBB '

Eq. 5.23

F0 = maximum force initially exerted on the beam by the impacting liquid


VAA = initial shear force at critical section AA
VBB = initial shear force at critical section BB
Because of the symmetry of the beam and the loading with respect to midspan,
the initial shear forces at the two critical sections are equal. Thus, from Eq. 5.23:
Vmax 0 = F0 / 2

Eq. 5.24

Vmax0 = maximum initial shear force in the beam = VAA = VBB


F0 = maximum force initially exerted on the beam by the impacting liquid
Maximum initial shear forces in the beams in the impact tests (Vmax0) are listed in
Table 5.5. The maximum initial nominal shear stress in the beam (max0) is:

max 0 =

Vmax 0
bd

Eq. 5.25

b = beam width
d = effective depth of beam (0.75 in for all specimens)
Vmax0 = maximum initial shear force in the beam
Maximum initial nominal shear stresses in the beams in the impact tests (max0)
are listed in Table 5.5. The ratio max0/(fc)1/2 (fc is the compressive strength of the
concrete) for each specimen is listed in Table 5.5. This ratio ranged between 3
and 14.5. The wide range of maximum initial shear stress is consistent with the
different mechanisms of deformation and/or failure observed in the impact tests.
The state of the specimens after impact ranged from complete disintegration to

122
small permanent deflections accompanied by narrow cracks not exceeding 0.03
in. (section 4.3.3).
To study the relationship between initial shear demand and failure mechanism,
Table 5.6 compares the calculated value of max0/(fc)1/2 (Table 5.5) with the
observed damage in each specimen (Table 4.6). In general, beams with low
ratios max0/(fc)1/2 (lower than 6, for instance) did not show inclined cracking
and/or disintegration. These beams developed flexural mechanisms of failure
(section 4.3.3). Beams with high ratios of max0/(fc)1/2 (higher than 6) did show
inclined cracking and/or disintegration. This trend was observed for all the test
specimens, which had different dynamic properties and different ratios td/T (listed
in Table 5.6, with td calculated for the full impacting liquid body using Eq. 5.19).
This evidence suggests that the initial shear failure of beams in the impact tests
is related to shear strength and does not depend on vibration properties.

5.3.5. Dynamic Shear Strength of Test Specimens


A shear stress threshold above which the beams in the impact tests show initial
failure in shear can be established from the results listed in Table 5.6. In each
series at least one specimen had a mechanism of deformation/failure dominated
by flexure, and at least one specimen had a failure mechanism dominated by
inclined cracking. The idea is to establish, in each series, the range of initial
shear demand (represented with the parameter max0/(fc)1/2) within which a
change in the failure mechanism (from flexure to shear) occurred. This range of
variation of max0/(fc)1/2 is defined by lower and upper limits. The lower limit is the
maximum value of max0/(fc)1/2, in a given series, for which a specimen developed
a flexural mechanism (and did not experience initial inclined cracking and
disintegration). The upper limit is the minimum value of max0/(fc)1/2, in a given
series, for which a specimen had inclined cracking and/or disintegration. These
limits are listed in Table 5.7 for all series and plotted versus series number in

123

Figure 5.11. From this figure, 6 f 'c [psi] seems to be a reasonable shear stress
threshold above which beams in the impact tests showed initial failure in shear.
As noted before, maximum initial nominal shear stresses (max0) ranged between

3 f 'c and 14.5 f 'c [psi] in the impact tests (Table 5.6). No beams with max0
lower than 6 f 'c (11 beams in total) showed inclined cracking and/or
disintegration (Table 5.6). These beams developed flexural mechanisms of
failure. Eighteen out of twenty beams with max0 higher than 6 f 'c had inclined
cracking and/or disintegration (Table 5.6). This was observed for test specimens
that had different dynamic properties and different ratios of duration of impact
load (td) to calculated period of vibration (T). The inferred nominal dynamic shear
strength of the tests specimens, Vnd [lbf], is:

Vnd = 6 f 'c bd

Eq. 5.26

fc = compressive strength of concrete (psi)


b = beam width (in)
d = effective depth of beam (in)
The ratio of the suggested nominal dynamic shear strength (Eq. 5.26) to the
nominal static shear strength (Eq. 5.7) of the beams in the impact tests is 1.6.

5.3.6. Deflections of Specimens with Flexural Mechanisms of Failure


The dynamic response of specimens that had deformation/failure mechanisms
dominated by flexure in the impact tests can be studied using the idealizations
described in sections 5.3.2 and 5.3.3. Only the specimens for which deflection
measurements were obtained (Table 4.6) during the impact tests are considered.

124
The impact loads were idealized as shown in Figure 5.9, with the parameters F0
and td (listed in Table 5.4) calculated using Eq. 5.19-5.20.
Equivalent SDOF systems were used to idealize the beams. The mass factor
(KM) and the mass of the SDOF system (Me) were assumed to be constant and
equal to 0.37 and 0.37 times the mass of the beam, respectively (section 5.3.2).
Values of Me for the different specimens analyzed are listed in Table 5.3. The
SDOF systems used had elasto-plastic resistance functions (Figure 5.8) with
maximum static resistances Rm as listed in Table 5.3. The resistances of the
beams under impact load can be higher than under static load because of
possible increases in the yield stress (and strength) of the steel reinforcement at
high strain rates. A simple way to account for this increase in beam strength is to
multiply the maximum static resistance by a dynamic increase factor (DIF) to
obtain the maximum dynamic resistance, Rmd (Figure 5.8):

Rmd = DIF Rm

Eq. 5.27

DIF = dynamic increase factor


Rm = maximum static resistance of the SDOF system
DIF depends on the strain rate in the steel reinforcement ( s ). According to
Malvar and Crawford (1998), a review of loading-rate effects on concrete and
reinforcing steel (Fu et. al., 1991) indicates that the modulus of elasticity remains
nearly constant. Thus, the effective stiffness of the beam is assumed to remain
constant and the values of ke for the equivalent SDOF systems (Table 5.3) were
estimated from static load-deflection curves (Figure 4.4 to Figure 4.14). The
yield deflection (of the SDOF system) under dynamic load, yd (Figure 5.8), is:

125

yd =

Rm d
R
= DIF m = DIF y
ke
ke

Eq. 5.28

ke = stiffness of equivalent SDOF system


y = yield deflection (of the SDOF system) under static load
The parameters Rmd and yd depend on DIF. Estimating DIF is not a simple task
because the available data on dynamic properties of steel reinforcement is
limited and there is uncertainty in the magnitude of s . Two simple approximate
methods are proposed below to estimate the values of s in the impact tests.
Method 1
Consider the elasto-plastic static resistance function of the equivalent SDOF
system (Figure 5.8). Yield deflections (y) are listed in Table 5.3 for all series.
Assume that the static yield stress of the reinforcement in tension in a specimen
occurs simultaneously at midspan and near the supports (the beams have fixed
ends and are loaded at midspan) at a deflection equal to y. The time at which y
was reached (ty) was obtained from the measured midspan deflection history for
each specimen (Figure 4.75-Figure 4.82) Values of ty are listed in Table 5.8. At
t = ty the strain in the reinforcement is:

y =

fy
Es

Eq. 5.29

y = static yield strain of reinforcement


fy = static yield stress of reinforcement
Es = modulus of elasticity of reinforcement
Table 5.8 contains values of y calculated using Eq. 5.29 and measured values of
fy and Es (Table A.5). The strain rate in the reinforcement ( s ) is estimated as:

126

s =

y
t y

Eq. 5.30

y = static yield strain of reinforcement


ty = time at which the static yield deflection (y) is reached
Estimates of s (Table 5.8) range approximately between 10-30 s-1 for specimens
in series 1, 3, 5, and between 3-21 s-1 for specimens in series 7-11.
Method 2
Assume that the specimen responds in the plastic range after it reaches its
maximum midspan deflection, max (at a time tmax). max and tmax are obtained
from the measured midspan deflection histories of the specimens (Figure 4.75 to
Figure 4.82) and are listed in Table 5.8. In the plastic range of response, three
plastic hinges form in the beam: one at each end and one at midspan. When max
is reached, the plastic-hinge rotations (max) at the beam ends are:

max =

max
L/2

Eq. 5.31

max = maximum midspan deflection


L = clear span
Assume that beam curvature is concentrated in the plastic-hinge region, and that
this region spreads along the span over a length equal to the beam depth, h. If
the curvature is assumed to be constant within the plastic hinge:

max = max h
max = plastic-hinge rotation associated with max

max = curvature in the plastic hinge associated with max


h = beam depth

Eq. 5.32

127
From Eq. 5.31-5.32 and considering that h = 1 in. for all specimens:

max =

2 max
in 1
L

Using Eq. 5.31 and Eq. 5.33, max and


(Table 5.8). From

Eq. 5.33

max are obtained for each specimen

max , the depth of the neutral axis at a beam section in the

plastic hinge (cmax) can be obtained using FLECHA (Pujol, 2001). Assuming
linear distribution of strains across the depth of the beam:

max =

s max

d cmax

Eq. 5.34

max = curvature in the plastic hinge associated with max


smax = strain of steel reinforcement associated with max
d = effective depth of beam section
cmax = depth of neutral axis at beam section in the plastic hinge
associated with max [in]
max = maximum midspan deflection
From Eq. 5.34 and considering that d = 0.75 in. for all specimens:

s max = (0.75 cmax ) max


where cmax and

Eq. 5.35

max have in. and 1/in. units, respectively. Values of cmax and max

are listed in Table 5.8. The strain rate in the reinforcement ( s ) is estimated as:

128

s =

s max
t max

Eq. 5.36

tmax = time at which max is reached


Estimates of s (Table 5.8) range approximately between 8-24 s-1. Method 2 is
expected to overestimate s in beams with small midspan deflections.
Methods 1 and 2 give estimates of

s of the same order of magnitude. The

differences in results do not seem large given the major approximations and
assumptions made. Combining the results from both methods, lower and upper
bounds for s can be obtained (Table 5.8). For specimens in series 1, 3, 5,

s ranges approximately between 8-30 s-1. For specimens in series 7-11,


s ranges approximately between 3-24 s-1.
Malvar and Crawford (1998) reviewed experimental data on the effects of high
strain rates on the properties of steel reinforcing bars. These data are limited for
strain rates higher than 2 s-1. Within the ranges of strain rates ( s ) estimated for
the impact tests (Table 5.8), the data show that the dynamic increase factor (DIF)
can be as low as 1.3 and as high as 1.7 for yield stresses (fy) between 40-80 ksi.
The data also show that as fy increases, DIF decreases. For steel reinforcing
bars with fy ranging between 42-103 ksi, a relationship that gives DIF as a
function of s and fy was proposed:


DIF = s4
10

Eq. 5.37

where:

= fy = 0.074 0.040

fy
60

Eq. 5.38

129
And:

s = strain rate in steel reinforcement [1/s]


fy = static yield stress of steel reinforcement [ksi]
The relationship proposed by Malvar and Crawford (1998), i.e., Eq. 5.37-5.38,
gives values of and DIF close to 0 and 1, respectively, for fy = 103 ksi. Because
it was observed that DIF decreases with increasing fy, DIF is expected to be
close to 1 for fy = 120 ksi. Thus, DIF = 1 is selected for test specimens in series
1, 3, and 5 (Table 4.1). For test specimens in series 7-11 (fy = 50 ksi, Table 4.1;
3 s-1 < s < 24 s-1, Table 5.8), Eq. 5.37-5.38 give = 0.041 and DIF ranging
approximately from 1.5 to 1.6. For simplicity, DIF = 1.5 is selected for series 7-11.
The parameters that define the dynamic resistance functions of the equivalent
SDOF systems, Rmd and yd (Figure 5.8), are obtained from the selected values
of DIF using Eq. 5.27-5.28 (Table 5.9).
The -method proposed by Newmark (1959) was used for dynamic analyses of
the equivalent SDOF systems subjected to the idealized impact loads. The
analyses were done using an algorithm implemented in Mathcad (PTC, 2007)
and presented in Appendix B. Calculated midspan deflection histories of the test
specimens (from the SDOF analyses) are shown in Figure 5.12 to Figure 5.21.
Calculated maximum midspan deflections, (max)calc., are listed and compared
with measured maximum midspan deflections (max) in Table 5.10. The times at
which max and (max)calc. are reached (tmax and t(max)calc., respectively) are also
listed in Table 5.10. The mean values of the ratios (max)calc./max and
t(max)calc./tmax are 1.11 and 0.78.
Comparisons of the measured (Figure 4.75 to Figure 4.82) and calculated
(Figure 5.12 to Figure 5.21) midspan deflection histories show that measured
and calculated dynamic responses have similar general behaviors but different
maximum and permanent deflections. All beams have increasing midspan

130
deflection up to a maximum value (max and (max)calc.). After the maximum
midspan deflection is reached, rebound and small-amplitude oscillation about
the permanent midspan deflection of the beam occur. Periods of vibration in this
range of response are shorter in the calculations than in the measurements. This
is consistent with the fact that as the beams are impacted by the liquid bodies,
wide cracks form and concrete spalls, which cause lengthening of the effective
periods of the beams. This softening was not considered in the analyses of
dynamic response of the equivalent SDOF systems (Appendix B). Thus, in the
non-linear range of response the SDOF systems used in the calculations are
stiffer than the test specimens. This is consistent with the fact that the SDOF
systems oscillate about their final positions (permanent deflections) with periods
of vibration T similar to those listed in Table 5.3. The beam rebound after max is
reached is smaller in the calculations than in the measurements.
In general, (max)calc. is larger than max. Even though the mean value of
(max)calc./max is 1.11, the error in (max)calc. was as high as 50% for specimens
S1_T1, S5_T4, S7_T4, S9_T1, and 70% for specimen S5_T3. In general,
t(max)calc. is shorter than tmax. Even though the mean value of t(max)calc./tmax is
0.78, the error in t(max)calc. is approximately 60% for specimens S3_T1 and
S9_T1 (for which the error in (max)calc. was between 40-50%). The errors in
(max)calc. and t(max)calc. are not intolerable in view of the uncertainties involved in
the calculation process regarding impact loads and beam resistances.
The idealization of the impact load used (Figure 5.9, with F0 and td calculated
using Eq. 5.19-5.20) implies that no beam deformation occurs within the duration
of the load (section 5.3.3). This can only be true for impact on a rigid barrier or for
impulsive impact load. The small-scale RC beams in the impact tests clearly
cannot be regarded as rigid barriers, and the impact loads are not impulsive
(section 5.3.3). The specimens analyzed in this section had idealized impact load
durations in the range 1-2 ms (Table 5.10). Within the load durations, significant

131
midspan deflections (ranging between 20-70% of max) occur in the specimens
(Figure 4.75-Figure 4.82). This means that the beam deforms and moves within
the duration of the load and, therefore, the relative velocity between the beam
and the liquid body is lower than the projectile velocity. The use of Eq. 5.20
(intended to give an estimate of initial shear demand) to estimate the impact
force at all times to obtain overall dynamic response of the test specimens is
questionable. The interaction between the liquid and the beam is complex and
should be studied incrementally for better results.
There is also uncertainty in the calculations of midspan deflections related to the
estimation of dynamic resistance of the specimens. It was assumed that the
beams had dynamic resistance functions as shown in Figure 5.8. The maximum
dynamic resistance Rmd was obtained using a dynamic increase factor (DIF) that
depends on the strain rate in the reinforcement ( s ). The strain rates are not
easy to compute and were estimated using approximate methods. Relevant
experimental data for the selection of DIF is limited.
Calculated Beam Deflections and Strain Rate in the Reinforcement using Finite
Element Analysis (FEA)
FEA was used to simulate the response of test specimens to fluid impact. One
goal of the analyses was to obtain results for strain rates in the reinforcement
that could be compared with the values estimated using Methods 1-2 above.
Another goal was to test the reliability of the FEA software used.
The FEA program LS-DYNA (LSTC, 2005) was used to simulate the response of
specimens S3_T1, S7_T1, S8_T1 (Table 4.6) to impact of a liquid body. A typical
finite-element (FE) model built in LS-DYNA is shown in Figure 5.22a. The smallscale RC beams were represented with numerical models that used constantstress solid hexahedral elements for the concrete and beam elements for the
steel reinforcement. Bond between the concrete and the reinforcement was

132
modeled using the *CONSTRAINED_LAGRANGE_IN_SOLID command in LSDYNA (LSTC, 2003). Aluminum containers used in the impact tests (with a mass
of approximately 0.02 lb) were modeled with shell elements. The dynamic
behavior of the impacting liquid body (water) was modeled using smoothparticle hydrodynamics (SPH; LSTC, 2003). The liquid body (with a mass of
approximately 0.4 lb) was set into motion using the *INITIAL_VELOCITY_
GENERATION command in LS-DYNA (LSTC, 2003). The velocities of the
impacting bodies (projectile velocity, vp) are listed in Table 4.6. Contact between
the impacting liquid body and the beam was modeled using the automatic
nodes-to-surface contact algorithm in LS-DYNA (LSTC, 2003).
The beam models had their ends restrained against rotation. This condition was
achieved by using two steel plates at each end (one plate on each side of the
beam) that were in direct contact with the beam (Figure 5.22b-c). These plates
had all their nodes restrained against displacement and rotation. Contact
between the support plates and the beam was modeled using the automatic
surface-to-surface contact algorithm in LS-DYNA (LSTC, 2003). The support
plates also restrained the ends of the beam against displacement in the direction
of travel of the impacting body (Figure 5.22b). Single-node restraints were used
at the ends of the beam to prevent displacements in the horizontal direction
perpendicular to the direction of impact (Figure 5.22b-c). Vertical displacement
was unrestrained at both ends of the beam.
Small-scale concrete was represented using LS-DYNA Material Type 84, which is
a smeared crack concrete model that includes strain rate effects (LSTC, 2003).
Rate effects on concrete were accounted for using the formulation proposed by
Malvar and Ross (1998). Steel and aluminum were represented using LS-DYNA
Material Type 3, which is suited to model isotropic and kinematic hardening
plasticity with the option of including rate effects (LSTC, 2003). Rate effects on
steel were accounted for using the formulation proposed by Malvar (1998). Water

133
was represented with LS-DYNA Material Type 9 (Null Material), which has no
yield strength and behaves in a fluid-like manner (LSTC, 2003). The following
material properties were used in the LS-DYNA models:
- Concrete:
Compressive strength (fc): As listed in Table 4.1
Modulus of elasticity (Ec): As listed in Table A.1
Poissons ratio (c): 0.15
Mass density (c): 2.2 x 10-4 lbf-s2/in4 (2400 kg/m3)
- Steel:
Yield stress (of reinforcement, fy): As listed in Table 4.1
Yield stress (of support plates, Fy): 50 ksi (340 MPa)
Modulus of elasticity (Es): 29,000 ksi (200 GPa)
Poissons ratio (c): 0.3
Mass density (s): 7.3 x 10-4 lbf-s2/in4 (7800 kg/m3)
- Aluminum:
Yield stress (Fy): 50 ksi (275 MPa)
Modulus of elasticity (Ea): 10,700 ksi (69 GPa)
Poissons ratio (a): 0.3
Mass density (a): 2.6 x 10-4 lbf-s2/in4 (2700 kg/m3)
- Water:
Mass density (w): 9.4 x 10-5 lbf-s2/in4 (1000 kg/m3)
The failure criterion used for reinforcing steel and for aluminum was based on
plastic (true) strain. In this criterion, an element is assumed to fail and is removed
from the model when the plastic strain exceeds a limit. This limit was fixed at
30% and 20% for steel and aluminum, respectively. Erosion of concrete was
assumed to occur when the maximum principal (true) strain exceeded 10%.

134
A typical rendering of the LS_DYNA simulations of fluid impact on test specimens
S3_T1, S7_T1, S8_T1 is shown in Figure 5.23. This graph shows the condition of
specimen S3_T1 at the time of maximum midspan deflection (7.5 ms). Midspan
deflection histories computed using FEA are shown and compared with
measured midspan deflection histories in Figure 5.24 to Figure 5.26. Maximum
midspan deflections from FEA results ((max)FEA) are listed and compared with
measured deflections (max) and deflections calculated using SDOF analysis
((max)calc.) in Table 5.11. The mean values of the ratios (max)FEA/max and
(max)FEA/(max)calc. are 0.73 and 0.9, respectively. The times at which max,
(max)calc., and (max)FEA are reached (tmax, t(max)calc., and t(max)FEA, respectively)
are listed in Table 5.11. The mean values of the ratios t(max)FEA/tmax and
t(max)FEA/t(max)calc. are 0.92 and 1.58, respectively.
From the mentioned results, it is observed that LS-DYNA produces estimates of
response similar to the measurements, but underestimates maximum midspan
deflections. A model as simple as an SDOF system gave estimates of maximum
midspan deflections comparable to the ones obtained using FEA, as implied by
the mean value of 0.9 for (max)FEA/(max)calc.. FEA was more time-consuming.
The time required to develop an FE model in LS-DYNA and define the input data
to be used in a given simulation was approximately 2 hrs. The LS-DYNA
simulations required computation times of up to 10 hrs using an AMD Opteron
computer with a single-core Processor 250 and 4 GB of system memory.
Histories of strain in the reinforcement in tension at midspan and near a support
are shown in Figure 5.27 to Figure 5.29. In these curves, the slope at a given
time gives an estimate of the instantaneous strain rate in the reinforcement ( s ).
In general, s increases at a higher rate at midspan than near a support.
Considering the steepest initial slopes in Figure 5.27 to Figure 5.29, s at
midspan is estimated to be as high as 22 s-1, 20 s-1, and 13 s-1 for specimens

135
S3_T1, S7_T1, S8_T1, respectively. Similarly, s near a support is estimated to
be as high as 25 s-1 for specimen S3_T1 and 6 s-1 for specimens S7_T1 and
S8_T1. These estimates of s are consistent with the values listed in Table 5.8.

136

Table 5.1 Calculated vs. Measured Initial Stiffnesses and Cracking Loads of Beams in the Static Tests
Series

Gross
Moment
of Inertia,
Ig
4
(in )

Clear
Span,
L
(in)

Compressive
Strength of
Concrete,
(1)
fc
(psi)

Modulus of
Elasticity of
Concrete,
(2)
Ec
(ksi)

Modulus of
Rupture of
Concrete,
(3)
fr
(psi)

Calculated
Cracking
Moment,
(Mcr)calc.
(lbf-in)

Calculated
Cracking
Load,
(Pcr)calc.
(lbf)

Calculated
Initial
Stiffness,
(k0)calc.
(kip/in)

Measured
Cracking
Load,
Pcr
(lbf)

Measured
Initial
Stiffness,
k0
(kip/in)

(Pcr)calc./
Pcr

(k0)calc./
k0

0.17

12

8800

5350

1130

380

250

99

280

100

0.89

0.99

0.17

12

8500*

5250

1110

370

250

97

280

86

0.88

1.13

0.33

12

7200

4840

1020

680

450

179

540

160

0.84

1.12

0.33

12

8200

5160

1090

720

480

191

540

170

0.89

1.12

0.50

12

8700

5320

1120

1120

750

295

500

145

1.49

2.04

0.50

12

9700

5610

1180

1180

790

312

750

180

1.05

1.73

0.17

12

6300

4650**

950

320

210

86

200

75

1.06

1.15

0.17

12

5600

4230**

900

300

200

78

200

50

1.00

1.57

0.33

12

5400

3950**

880

590

390

146

350

100

1.12

1.46

10

0.33

12

6700

4190**

980

660

440

155

500

115

0.87

1.35

11

0.33

7400

4480**

1030

690

690

559

780

300

0.88

1.86

Mean

1.00

1.41

Standard Deviation

0.19

0.35

(1)

Estimated compressive strength of concrete at day of static test (from Figure A.1-Figure A.4).

(2)

Estimated modulus of elasticity of concrete at day of static test (using Eq. 5.4, i.e., Ec = 57,000*(fc) ).

(3)

Estimated modulus of rupture of concrete at day of static test (using fr = 12*(fc) , section 5.2.2).

1/2

1/2

** Estimates of Ec using Eq. 5.4 were larger than measured values of Ec after impact tests. The measured values of Ec were used to calculate k0.

136

* Compressive strength (fc) was not measured in series 2 because cylinders were defective (section A.2.1). The average of fc in series 1, 3-6 was used.

137

Table 5.2 Calculated vs. Measured Peak Loads and Maximum Nominal Shear Stresses in Beams in the Static Tests
Beam
Width,
b
(in)

Clear
Span,
L
(in)

Compressive
Strength of
Concrete,
(1)
fc
(psi)

Calc.
Limiting
Moment,
(Mmax)calc.
(lbf-in)

Calc.
Peak
Load,
(Pmax)calc.
(lbf)

Measured
Peak
Load,
Pmax
(lbf)

(Pmax)calc./
Pmax

12

8800

810

540

740

12

8500*

1350

900

12

7200

1540

12

8200

12

Series

(1)

Calc.
Limiting
Moment

Calc.
Peak
Load

a
(Pmax )calc./

Max.
Shear
Force,
Vmax
(lbf)

Max.
Nominal
Shear
Stress,
max
(psi)

max/

(axial force),

(axial force),

Pmax

0.73

1190

790

1.07

370

250

2.6

1060

0.85

1710

1140

1.07

530

350

3.8

1030

1270

0.81

1920

1280

1.01

635

210

2.5

2690

1800

2140

0.84

3050

2030

0.95

1070

360

3.9

8700

2420

1610

1740

0.93

2800

1870

1.07

870

190

2.1

12

9700

4180

2790

2960

0.94

4540

3030

1.02

1480

330

3.3

12

6300

370

250

350

0.70

740

490

1.41

175

120

1.5

12

5600

660

440

580

0.76

1000

670

1.15

290

190

2.6

12

5400

740

490

820

0.60

1190

790

0.97

410

140

1.9

10

12

6700

1390

920

1280

0.72

1740

1160

0.90

640

210

2.6

11

7400

820

820

1370

0.60

1240

1240

0.91

685

230

2.7

(Mmax )calc.
(lbf-in)

(Pmax )calc.
(lbf)

Mean

0.77

Mean

1.05

Standard Deviation

0.12

Standard Deviation

0.14

(fc)

1/2

Estimated compressive strength of concrete at day of static test (from Figure A.1-Figure A.4).

* Compressive strength (fc) was not measured in series 2 because cylinders were defective (section A.2.1). The average of fc in series 1, 3-6 was used.

137

138

Table 5.3 Dynamic Properties of Specimens and Equivalent SDOF Systems


Mass of Equivalent
SDOF System, Me
2
(lbf-s /in)

Stiffness of
Equivalent
SDOF System,
ke
(kip/in)

Max. Resistance
of Equivalent
SDOF System,
Rm
(lbf)

Yield
Deflection
of Equivalent
SDOF System,
y
(in)

Calculated
Effective
Period of
Vibration,
T
(ms)

Calculated
Uncracked
Period of
Vibration,
Tuncr.
(ms)

m/Me

Linear
Range

Plastic
Range

Measured
Initial Beam
Stiffness,
k0
(kip/in)

0.0055

0.0020

0.0018

100

42.9

700

0.016

1.4

0.9

0.51

0.0109

0.0040

0.0036

160

83.3

900

0.011

1.4

1.0

0.25

0.0164

0.0061

0.0054

145

100

1250

0.013

1.5

1.3

0.17

0.0055

0.0020

0.0018

75

47.5

300

0.006

1.3

1.0

0.51

0.0055

0.0020

0.0018

50

40

550

0.014

1.4

1.3

0.51

0.0109

0.0040

0.0036

100

70

550

0.008

1.5

1.3

0.25

10

0.0109

0.0040

0.0036

115

83.3

1100

0.013

1.4

1.2

0.25

11

0.0073

0.0027

0.0024

300

160

1200

0.008

0.8

0.6

0.38

Series

Beam
Mass,
M
2
(lbf-s /in)

138

139

Table 5.4 Idealized Impact Loads and Kinetic Energies

Specimen

S1_T1
S1_T2
S1_T3
S3_T1
S3_T2
S3_T3
S3_T4
S5_T1
S5_T2
S5_T3
S5_T4
S7_T1
S7_T2
S7_T3
S7_T4
S8_T1
S8_T2
S8_T3
S8_T4
S9_T1
S9_T2
S9_T3
S9_T4
S10_T1
S10_T2
S10_T3
S10_T4
S11_T1
S11_T2
S11_T3
S11_T4

Calculated
Effective
Period of
Vibration,
T
(ms)

Impact
Velocity,
vp
(in/s)

Duration of
Idealized
Impact Load,
td
(ms)

Magnitude
of Idealized
Impact Load,
F0
(lbf)

Kinetic
Energy of
Impacting
Body,
KEp
(lbf-in)

Kinetic
Energy
Transferred
to Beam,
KE0
(lbf-in)

td/T

1.4
1.4
1.4
1.4
1.4
1.4
1.4
1.5
1.5
1.5
1.5
1.3
1.3
1.3
1.3
1.4
1.4
1.4
1.4
1.5
1.5
1.5
1.5
1.4
1.4
1.4
1.4
0.8
0.8
0.8
0.8

1860
3370
3670
3430
4150
3950
4390
5560
5770
3260
2950
2210
2590
3230
2930
1840
2630
2280
3050
2570
3260
3150
4150
2780
3740
3460
4210
3020
3780
4550
5100

1.9
1.0
1.0
1.0
0.8
0.9
0.8
0.6
0.6
1.1
1.2
1.6
1.4
1.1
1.2
1.9
1.3
1.5
1.2
1.4
1.1
1.1
0.8
1.3
0.9
1.0
0.8
1.2
0.9
0.8
0.7

1020
3340
3960
3460
5060
4580
5660
9080
9780
3120
2560
1440
1970
3060
2520
990
2030
1530
2730
1940
3120
2920
5060
2270
4110
3520
5210
2680
4200
6080
7640

1790
5860
6950
6070
8890
8050
9950
15960
17180
5480
4490
2520
3460
5380
4430
1750
3570
2680
4800
3410
5480
5120
8890
3990
7220
6180
9150
4710
7380
10690
13420

910
2990
3540
1550
2270
2050
2540
2710
2920
930
760
1280
1760
2750
2260
890
1820
1370
2450
870
1400
1310
2270
1020
1840
1580
2330
1800
2820
4090
5130

1.38
0.76
0.70
0.74
0.61
0.64
0.58
0.41
0.39
0.70
0.77
1.23
1.05
0.84
0.92
1.35
0.95
1.09
0.82
0.90
0.71
0.74
0.56
0.91
0.68
0.73
0.60
1.43
1.14
0.95
0.84

140
Table 5.5 Maximum Initial Shear in Beams in the Impact Tests

Specimen

Section
Width,

(in)

S1_T1
S1_T2
S1_T3
S3_T1
S3_T2
S3_T3
S3_T4
S5_T1
S5_T2
S5_T3
S5_T4
S7_T1
S7_T2
S7_T3
S7_T4
S8_T1
S8_T2
S8_T3
S8_T4
S9_T1
S9_T2
S9_T3
S9_T4
S10_T1
S10_T2
S10_T3
S10_T4
S11_T1
S11_T2
S11_T3
S11_T4
(1)

2
2
2
4
4
4
4
6
6
6
6
2
2
2
2
2
2
2
2
4
4
4
4
4
4
4
4
4
4
4
4

Compressive
Strength of
Concrete,
(1)

fc

(psi)

9200
9200
9200
7630
7630
7630
7630
9120
9120
9120
9120
7040
7040
7040
7040
6330
6330
6330
6330
6390
6390
6390
6390
7690
7690
7690
7690
7720
7720
7720
7720

Impact
Velocity,

vp

(in/s)

1860
3370
3670
3430
4150
3950
4390
5560
5770
3260
2950
2210
2590
3230
2930
1840
2630
2280
3050
2570
3260
3150
4150
2780
3740
3460
4210
3020
3780
4550
5100

Maximum
Initial
Impact Force,

F0

(lbf)

1020
3340
3960
3460
5060
4580
5660
9080
9780
3120
2560
1440
1970
3060
2520
990
2030
1530
2730
1940
3120
2920
5060
2270
4110
3520
5210
2680
4200
6080
7640

Maximum
Initial
Shear
Force,

Maximum
Initial
Nominal
Shear Stress,

(lbf)

(psi)

Vmax0
510
1670
1980
1730
2530
2290
2830
4540
4890
1560
1280
720
985
1530
1260
495
1015
765
1365
970
1560
1460
2530
1135
2055
1760
2605
1340
2100
3040
3820

max0/(fc)

max0

340
1110
1320
580
840
760
940
1010
1090
350
280
480
660
1020
840
330
680
510
910
320
520
490
840
380
680
590
870
450
700
1010
1270

Measured compressive strength of concrete at age of impact tests (Table A.1 and Table A.2).

3.5
11.6
13.7
6.6
9.7
8.7
10.8
10.6
11.4
3.6
3.0
5.7
7.8
12.2
10.0
4.2
8.5
6.4
11.4
4.0
6.5
6.1
10.5
4.3
7.8
6.7
9.9
5.1
8.0
11.5
14.5

1/2

141
Table 5.6 Initial Shear Demand vs. Failure Mechanism in the Impact Tests
Specimen
S1_T1
S1_T2
S1_T3
S3_T1
S3_T2
S3_T3
S3_T4
S5_T1
S5_T2
S5_T3
S5_T4
S7_T1
S7_T2
S7_T3
S7_T4
S8_T1
S8_T2
S8_T3
S8_T4
S9_T1
S9_T2
S9_T3
S9_T4
S10_T1
S10_T2
S10_T3
S10_T4
S11_T1
S11_T2
S11_T3
S11_T4

Impact
Velocity,
vp
(in/s)

td/T

1860
3370
3670
3430
4150
3950
4390
5560
5770
3260
2950
2210
2590
3230
2930
1840
2630
2280
3050
2570
3260
3150
4150
2780
3740
3460
4210
3020
3780
4550
5100

1.38
0.76
0.70
0.74
0.61
0.64
0.58
0.41
0.39
0.70
0.77
1.23
1.05
0.84
0.92
1.35
0.95
1.09
0.82
0.90
0.71
0.74
0.56
0.91
0.68
0.73
0.60
1.43
1.14
0.95
0.84

max0/(fc)
3.5
11.6
13.7
6.6
9.7
8.7
10.8
10.6
11.4
3.6
3.0
5.7
7.8
12.2
10.0
4.2
8.5
6.4
11.4
4.0
6.5
6.1
10.5
4.3
7.8
6.7
9.9
5.1
8.0
11.5
14.5

1/2

Observed Damage
Disintegration
of Beam
No
Yes
Yes
No
Yes
No
Yes
Yes
Yes
No
No
No
No
Yes
No
No
No
No
Yes
No
Yes
No
Yes
No
No
No
Yes
No
Yes
Yes
Yes

Inclined
Cracking
No*
Yes
Yes
No*
Yes
No*
Yes
Yes
Yes
No*
No*
No*
No*
Yes
No*
No*
Yes
Yes
Yes
No*
Yes
No*
Yes
No*
Yes
Yes
Yes
No*
Yes
Yes
Yes

* Mechanism of deformation and/or failure dominated by flexure (section 4.3.3).

142
Table 5.7 Range of Variation of Initial Shear Demand for Change in the Failure
Mode of Beams in the Impact Tests
Series

Lower Limit of Upper Limit of


max0/(fc)1/2
max0/(fc)1/2

3.5

11.6

8.7

9.7

3.6

10.6

10.0

12.2

4.2

6.4

6.1

6.5

10

4.3

6.7

11

5.1

8.0

143

Table 5.8 Estimated Strain Rates in the Steel Reinforcement of Beams in the Impact Tests

Specimen

S1_T1
S3_T1
S5_T3
S5_T4
S7_T1
S7_T4
S8_T1
S9_T1
S10_T1
S11_T1

Impact
Velocity,
vp
(in/s)

1860
3430
3260
2950
2210
2930
1840
2570
2780
3020

Static
Yield
Deflection,
y
(in)

Time at
which
y is
reached,
ty
(ms)

Maximum
Midspan
Deflection,
max
(in)

Time at
which
max is
reached,
tmax
(ms)

Static
Yield
Strain,
y

0.016
0.011
0.013
0.013
0.006
0.006
0.014
0.008
0.013
0.008

0.37
0.12
0.19
0.2
0.08
0.11
0.27
0.28
0.5
0.23

0.29
1.6
0.24
0.22
1.2
1.5
0.29
1.2
0.32
0.39

2.8
7.5
2.0
2.0
5.3
5.1
2.8
6.7
2.5
2.7

0.0037
0.0037
0.0037
0.0037
0.0017
0.0017
0.0017
0.0017
0.0017
0.0017

Strain
Rate,

(1)

10.0
30.8
19.4
18.5
20.7
15.1
6.1
5.9
3.3
7.2

Plastic
Hinge
Rotation
at max,
max
(rad)

Curvature
at Plastic
Hinge
at max,

0.048
0.267
0.040
0.037
0.192
0.250
0.048
0.200
0.053
0.098

Strain
at
max,
smax

Strain
Rate,

(1/in)

Depth of
Neutral
Axis
at max,
cmax
(in)

0.048
0.267
0.040
0.037
0.192
0.250
0.048
0.200
0.053
0.098

0.27
0.26
0.27
0.27
0.26
0.26
0.26
0.26
0.26
0.25

0.023
0.131
0.019
0.018
0.094
0.123
0.024
0.098
0.026
0.049

8.3
17.4
9.4
8.7
17.7
23.9
8.4
14.5
10.2
17.7

1.2
1.8
2.1
2.1
1.2
0.6
0.7
0.4
0.3
0.4

Mean

1.08

Standard Deviation

0.70

max

(2)

s (1) /
s ( 2)

143

144

Table 5.9 Dynamic Resistance Functions of Equivalent SDOF Systems


Series

Stiffness,
ke
(kip/in)

Maximum
Static
Resistance,
Rm
(lbf)

Static
Yield
Deflection,
y
(in)

Dynamic
Increase
Factor,
DIF

Maximum
Dynamic
Resistance,
Rmd
(lbf)

Dynamic
Yield
Deflection,
yd
(in)

42.9

700

0.016

1.0

700

0.016

83.3

900

0.011

1.0

900

0.011

100

1250

0.013

1.0

1250

0.013

47.5

300

0.006

1.5

450

0.009

40

550

0.014

1.5

825

0.021

70

550

0.008

1.5

825

0.012

10

83.3

1100

0.013

1.5

1650

0.020

11

160

1200

0.008

1.5

1800

0.011

Table 5.10 Calculated vs. Measured Maximum Midspan Deflections in Beams in


the Impact Tests

Specimen

Impact
Velocity,
vp
(in/s)

Duration
of
Idealized
Impact
Load,
td
(ms)

Maximum
Midspan
Deflection,
max
(in)

Time at
which
max
is
reached,
tmax
(ms)

Calculated
Maximum
Midspan
Deflection,
(max)calc.
(in)

Time at
which
(max)calc.
is
reached,
t(max)calc.
(ms)

(max)calc./
max

t(max)calc./
tmax

S1_T1
S3_T1
S5_T3
S5_T4
S7_T1
S7_T4
S8_T1
S9_T1
S10_T1
S11_T1

1860
3430
3260
2950
2210
2930
1840
2570
2780
3020

1.9
1.0
1.1
1.2
1.6
1.2
1.9
1.4
1.3
1.2

0.29
1.6
0.24
0.22
1.2
1.5
0.29
1.2
0.32
0.39

2.8
7.5
2.0
2.0
5.3
5.1
2.8
6.7
2.5
2.7

0.43
0.91
0.41
0.32
1.3
2.2
0.30
0.60
0.26
0.36

2.6
3.1
2.5
2.3
3.8
4.2
2.3
2.9
1.8
1.7

1.48
0.57
1.71
1.45
1.13
1.47
1.03
0.50
0.81
0.92

0.93
0.41
1.22
1.14
0.71
0.82
0.82
0.43
0.71
0.62

Mean

1.11

0.78

Standard Deviation

0.41

0.27

145

Table 5.11 Measured and Calculated (SDOF) Maximum Midspan Deflections vs. Results from FEA - Impact Tests

Specimen

Impact
Velocity,
vp
(in/s)

Maximum
Midspan
Deflection
(measured),
max
(in)

Time at
which
max
is
reached,
tmax
(ms)

Calculated
Maximum
Midspan
Deflection
(SDOF),
(max)calc.
(in)

Time at
which
(max)calc.
is reached,
t(max)calc.
(ms)

Calculated
Maximum
Midspan
Deflection
(FEA),
(max)FEA
(in)

Time at
which
(max)FEA
is reached,
t(max)FEA
(ms)

(max)FEA/
max

(max)FEA/
(max)calc.

t(max)FEA/
tmax

t(max)FEA/
t(max)calc.

S3_T1

3430

1.6

7.5

0.91

3.1

1.2

7.8

0.75

1.32

1.04

2.52

S7_T1

2210

1.2

5.3

1.3

3.8

0.62

4.0

0.52

0.48

0.75

1.05

S8_T1

1840

0.29

2.8

0.30

2.3

0.27

2.7

0.93

0.90

0.96

1.17

0.73

0.90

0.92

1.58

Mean

145

146

PL/8

P/2

L/4
L/2
L/4

PL/8
L/4
L/2
L/4
P/2

(a) Beam

(b) Shear Force (V) Diagram

PL/8
(c) Moment (M) Diagram

Figure 5.1 Shear Force and Moment Diagrams for Fixed-Fixed Beam with
Concentrated Load at Midspan

147

= 1%
= 0.5%

Figure 5.2 Calculated Moment-Curvature Relationships for Beams in Series 1-2

= 1%

= 0.5%

Figure 5.3 Calculated Moment-Curvature Relationships for Beams in Series 3-4

148

= 1%

= 0.5%

Figure 5.4 Calculated Moment-Curvature Relationships for Beams in Series 5-6

= 1%
= 0.5%

Figure 5.5 Calculated Moment-Curvature Relationships for Beams in Series 7-8

149

= 1%
= 0.5%
= 0.5%

Figure 5.6 Calculated Moment-Curvature Relationships for Beams in Series 9-11

M (mass),

k (stiffness)

Fe

ke
Me

(impact load)

(b) Equivalent SDOF System

(a) Test Beam

Figure 5.7 Representation of Test Specimen with an Equivalent SDOF System

150

Force

Rmd

Dynamic resistance function

Rm

Static resistance function

ke

y yd

Displacement

Figure 5.8 Elasto-Plastic Resistance Function of Equivalent SDOF System

Force
F0

td

Time

Figure 5.9 Idealized Impact Load

151

Segment of liquid
with w, A, m

Critical
section

vp

2 in

F0

A
1 in

(3.5 in)

A
1 in

L
lw

VAA
F0

(a) Impact of Small Segment


of Liquid on Test Beam

(finertia)ABBA 0
B

VBB

Critical
section

Impacting liquid body


with w, A, m

(c) Free Body Diagram


of Central Portion
of Test Beam

(b) Initial Impact Force and


Critical Sections for Shear

Figure 5.10 Initial Impact Loading and Beam Response

Initial Shear Demand, max0 /(f'c)1/2

14
12
10
8
6
4
Lower Limit (flexure)

Upper Limit (shear)


Assumed Dynamic Shear Strength

0
0

10

11

Test Series

Figure 5.11 Variation of Initial Shear Demand for Change in the Failure Mode
and Suggested Dynamic Shear Strength of Beams in the Impact Tests

Midspan Deflection (in)

152

Calculated
Maximum Measured

Time (s)

Midspan Deflection (in)

Figure 5.12 Calculated Midspan Deflection History Specimen S1_T1

Maximum Measured

Calculated

Time (s)

Figure 5.13 Calculated Midspan Deflection History Specimen S3_T1

Midspan Deflection (in)

153

Calculated
Maximum Measured

Time (s)

Midspan Deflection (in)

Figure 5.14 Calculated Midspan Deflection History Specimen S5_T3

Calculated
Maximum Measured

Time (s)

Figure 5.15 Calculated Midspan Deflection History Specimen S5_T4

Midspan Deflection (in)

154

Calculated
Maximum Measured

Time (s)

Figure 5.16 Calculated Midspan Deflection History Specimen S7_T1

Midspan Deflection (in)

Calculated

Maximum Measured

Time (s)

Figure 5.17 Calculated Midspan Deflection History Specimen S7_T4

Midspan Deflection (in)

155

Maximum
Measured

Calculated

Time (s)

Midspan Deflection (in)

Figure 5.18 Calculated Midspan Deflection History Specimen S8_T1

Maximum Measured

Calculated

Time (s)

Figure 5.19 Calculated Midspan Deflection History Specimen S9_T1

Midspan Deflection (in)

156

Maximum Measured
Calculated

Time (s)

Midspan Deflection (in)

Figure 5.20 Calculated Midspan Deflection History Specimen S10_T1

Maximum Measured
Calculated

Time (s)

Figure 5.21 Calculated Midspan Deflection History Specimen S11_T1

157

Support
plate

Container
filled with
water

vp

Specimen

(a) Overall View

(b) Side View

(c) Front View (specimen


and supports only)

Figure 5.22 Typical Finite Element Model to Calculate Response of Small-Scale


RC Beams to Fluid Impact using LS-DYNA

Support
plate

Container

Specimen
Water

(a) Overall View

(b) Front View

(c) Side View


(specimen and
supports only)

Figure 5.23 Typical Rendering of LS-DYNA Simulation of Fluid Impact on SmallScale RC Beams (Specimen S3_T1, 7.8 ms after impact)

158

1.8

1.5

Deflection (in)

1.2

0.9
Measured
Calculated using FEA

0.6

0.3

0.0
0.000

0.002

0.004

Time (s)

0.006

0.008

0.010

Figure 5.24 Calculated (FEA) Midspan Deflection History Specimen S3_T1

1.8

1.5
Measured
Calculated using FEA

Deflection (in)

1.2

0.9

0.6

0.3

0.0
0.000

0.002

0.004

Time (s)

0.006

0.008

0.010

Figure 5.25 Calculated (FEA) Midspan Deflection History Specimen S7_T1

159

1.8

1.5

Deflection (in)

1.2

0.9
Measured
Calculated using FEA

0.6

0.3

0.0
0.000

0.002

0.004

Time (s)

0.006

0.008

0.010

Figure 5.26 Calculated (FEA) Midspan Deflection History Specimen S8_T1

0.050

0.040

At Midspan

0.030

Strain

Near Support

0.020

0.010

0.000
0.000

0.002

0.004

Time (s)

0.006

0.008

0.010

Figure 5.27 Calculated (FEA) History of Strain in the Reinforcement in Tension


Specimen S3_T1

160

0.050

0.040

At Midspan

0.030

Strain

Near Support

0.020

0.010

0.000
0.000

0.002

0.004

Time (s)

0.006

0.008

0.010

Figure 5.28 Calculated (FEA) History of Strain in the Reinforcement in Tension


Specimen S7_T1

0.050

0.040

0.030

Strain

At Midspan
Near Support

0.020

0.010

0.000
0.000

0.002

0.004

Time (s)

0.006

0.008

0.010

Figure 5.29 Calculated (FEA) History of Strain in the Reinforcement in Tension


Specimen S8_T1

161

CHAPTER 6. ANALYSIS OF INITIAL RESPONSE OF BEAMS TO BLAST LOAD

6.1. Introduction
The response of small-scale reinforced concrete (RC) beams to concentrated
fluid impact load was discussed in chapter 4 and chapter 5. The initial response
of beams to uniformly distributed blast load will be discussed in this chapter. This
topic is relevant to the design of blast-resistant structural elements. For RC
elements the design procedure involves (McCann and Smith, 2007): (1)
Definition of the blast load, (2) adoption of a design criterion (e.g. maximum
displacement), (3) preliminary selection of element dimensions, materials and
reinforcement, (4) dynamic analysis of element response to the design blast load,
(5) comparison of the calculated response parameter with the adopted criterion
and adjustment of the element dimensions and reinforcement as needed to meet
the criterion, and (6) proper reinforcement detailing to ensure the elements have
sufficient toughness to reach the maximum deformations associated with the
adopted response limit.
In step (4) of the design procedure outlined above, it is common to use a SDOF
system to idealize the structural element analyzed. The stiffness, mass, and
loads used to define the SDOF system are selected so that its deflection is equal
to the maximum deflection of the element. Biggs (1964) developed a method to
compute the properties (stiffness, mass, load) of the equivalent SDOF system
by equating the kinetic and internal energies of the SDOF system and the
element it represents. The procedure requires an assumption about the deflected
shape of the structural element. Biggs suggested the use of the deflected shape
associated with static load.

162
Experiments (chapter 2 and chapter 4) and numerical results (chapter 3) have
shown that beams that fail in flexure under static load may fail in shear under
rapidly applied load. Shear failure limits the energy-dissipation capacity of a
beam, which is the key to the survival of the beam to blast load. If a designer is to
rely on high energy-dissipation capacity provided by a structural member
designed according to the procedure outlined above, the possibility of shear
failure during the initial phase of response of the element to blast load must be
addressed. Thus, the initial shear demand imposed by the blast load on a beam,
as well as the dynamic shear strength of the beam must be determined. The
focus of this chapter is the initial shear demand in beams subjected to blast load.
This demand is investigated using different approaches.
Several methods to compute shear demand in beams subjected to blast load are
available (Keenan, 1977; Murtha and Crawford, 1981; Ross and Krawinkler,
1985; Krauthammer, 1990). An objective look at these methods reveals lack of
consensus about the problem. Keenan (1977), and Murtha and Crawford (1981)
link initial shear demand to flexural strength. The Department of the Army (1990),
through the work summarized into technical manual TM5-1300, also suggests
computing maximum shear demand based on the ultimate flexural resistance of
the element (accounting for dynamic increase of flexural strength). Ross and
Krawinkler (1985), and Krauthammer (1990) do not link initial shear demand to
flexural strength. And all of these approaches rely heavily on empirical
assumptions or complex mathematical solutions. Two different methods to
estimate initial shear demand will be proposed in this chapter. Before this can be
done, the characteristics of blast load and the corresponding initial response of
beams must be discussed (section 6.2).
The first method proposed to estimate initial shear demand is based on the work
by Biggs (1964) to compute dynamic reactions and is discussed in section 6.4.
The dynamic response of a beam to blast load is obtained numerically using an

163
equivalent linear SDOF system. The second method proposed to estimate shear
demand uses finite differences to solve the differential equations of motion of a
linear Timoshenko beam subjected to uniform load (section 6.5). The results for
maximum shear obtained with the proposed methods are compared with results
from finite-element analysis (FEA). Initial shear demand and behavior of one-way
RC slabs tested under blast load (Slawson, 1984) are analyzed in section 6.6
using the methods presented in sections 6.4 and 6.5.

6.2. Characteristics of Blast Loads and Corresponding Initial Response of Beams


Blast load is the result of the interaction of a shock wave generated by an
explosion and the surface of a structural element. This load has very short
duration and high intensity, and is represented in terms of pressure (in excess of
the ambient air pressure) acting on the surface of the element. A typical
pressure-time history for a blast wave is shown in Figure 6.1a (Zhu et al. 2007).
At the time the shock wave first reaches a point on the surface of the element
(arrival time, ta), there is a rapid rise in pressure to a maximum value called p0.
The rise time is assumed to be zero in design. After the peak pressure p0 is
reached, the pressure decreases at a decreasing rate until a negative (below
ambient air pressure) peak is reached. From that point on, the pressure
increases again to ambient pressure as the shock wave moves away from the
element. The duration of the interval during which the pressure is positive at a
certain point (positive phase) is called dwell time, td. The interval during which the
pressure is negative (suction) at a certain point is referred to as negative phase.
The positive phase is more relevant to the response of a structural element than
the negative phase, which is typically ignored for design purposes. The unit
impulse imparted by the blast wave during the positive phase Is is:
Is =

ta +td

ta

p (t )dt

Eq. 6.1

164
Where:
t = time
ta = arrival time of blast wave
td = positive phase duration of blast wave (dwell time)
p = overpressure
The pressure resulting from a blast load can be idealized as a triangular pulse
with duration td and peak intensity p0 (Figure 6.1b). The unit impulse becomes:

Is = i =

p0 t d
2

Eq. 6.2

p0 = peak overpressure from blast load


td = blast pulse duration (dwell time)
The duration of the blast pulse is in the order of milliseconds. If this duration (td)
is short compared with the period of vibration of the structural element (T), the
blast load is referred to as being impulsive and its effect is idealized as an initial
velocity imparted to the structural element, v0:

v0 =

i
Me

Eq. 6.3

i = unit impulse of blast load


Me = effective mass of structural element
The mass is the only structural parameter that controls the initial motion of the
element caused by an impulsive load. It is common to assume that a load is
impulsive if td < 0.1T (Biggs, 1964). Assuming that a blast load is impulsive
simplifies the dynamic analysis. Thus, the response of a beam to blast load can

165
be determined by studying the free vibration of an equivalent SDOF oscillator
subjected to an initial velocity v0. From this analysis, the maximum displacement
that the element is likely to undergo can be determined. Maximum displacement
can also be obtained from energy balance considerations. The kinetic energy
imparted to the structural member by the blast load (assumed to be impulsive) is:

KE 0 =

i2
2M e

Eq. 6.4

where i is the unit impulse and Me is the effective mass of the element.
The kinetic energy imparted to the element (KE0) must be absorbed through
deformation. The required strain energy SE must be equal to KE0. Using Eq. 6.4,
and from the load-displacement relationship for the element (assumed to be
similar to the one for static load) the maximum deflection imposed by the blast
load is obtained. A beam must be designed to reach this deflection without losing
its structural integrity. Shear failure is associated with loss of integrity at smaller
deflections. As suggested in section 3.4, shear failure may occur if the dynamic
shear demand exceeds the capacity of the beam to resist shear. The shear
demand is the focus of the following sections.
The numerical results presented in chapter 3 showed that for beams that failed in
shear under blast load, the maximum shear occurred at the end section and
within the duration of the load td. Because the main concern of this chapter is the
maximum initial shear demand in a beam subjected to impulsive blast load, the
discussion in the following sections will be focused on response during the time
interval 0 t 0.1T . This range of response will be referred to as initial phase (or
initial response). The work presented in the following sections is done with a view
to the response of RC beams to blast load. Shear failures of RC elements are
brittle. Slawson (1984) showed experimentally that shear failures of RC slabs

166
subjected to blast load occurred at early times. Given the short times involved,
flexural deformations in the initial phase are expected to be small. These facts
(brittle nature of shear failure in concrete and small flexural deformations)
support the use of linear models to study the response of RC beams to blast load
in the initial phase. The methods to obtain initial shear demand (Vmax) presented
next deal with linear dynamic response. These methods are expected to give
upper-bound estimates of Vmax as non-linear behavior limits shear demand.

6.3. Classical Method to Compute Dynamic Reactions


Biggs (1964) showed that the maximum shear (or reaction force) V for a beam
under dynamic load can be estimated using the expression:
V (t ) = R(t ) + F (t )

Eq. 6.5

R = beam resistance
F = total applied load
, = factors that depend on the beam support conditions, load type, and
range of response (linear or non-linear)
Consider the case of a simply supported beam with uniform load (q), uniform
mass, and linear response (Figure 6.2a). The shape of the distribution of inertia
forces along the span is identical to the deflected shape of the beam (Figure
6.2b). The lateral acceleration y of a point in the beam at a distance x from a
support is:
y( x) = C ( x)

C = multiplier (varies with time)

(x) = beam deflected shape

Eq. 6.6

167
The deflected shape of a simply supported beam under uniform static load is:

( x) =

16 3
( L x 2 Lx 3 + x 4 )
4
5L

Eq. 6.7

x = distance from beam support


L= beam length (clear span)
Conventionally it is assumed that this deformed shape applies for static as well
as dynamic loads. Consider now the dynamic equilibrium of one half of the beam
as indicated in Figure 6.2c. By symmetry, the shear force at mid-span (S) must
be zero. For design purposes it is convenient to have an expression for V(t) that
does not include an inertia force term. Thus, rotational equilibrium about the
location of the resultant of the inertia forces ( x = 61L/192, measured from left
support) requires:

V (61L / 192) M m

1
F (61L / 192 L / 4) = 0
2

Eq. 6.8

V = end shear (support reaction)


L= beam length (clear span)
Mm = moment at mid-span
F = total applied load
Beam resistance R (variable with time) is defined as:

R(t ) =

8M m (t )
L

Mm = moment at mid-span
L= beam length (clear span)

Eq. 6.9

168
Substituting Eq. 6.9 into Eq. 6.8, and solving for V:
V (t ) = 0.39 R (t ) + 0.11F (t )

Eq. 6.10

R = beam resistance
F = total applied load
Eq. 6.10 applies to linear-elastic simply supported beams. Similarly, for linearelastic fixed-fixed beams with uniform load and uniform mass:
V (t ) = 0.36 R(t ) + 0.14 F (t )

Eq. 6.11

Following a similar procedure but assuming different deflected shapes,


expressions can be obtained for response in the plastic range, as well as for
beams with different types of loads and boundary conditions (Biggs, 1964). Eq.
6.10 and Eq. 6.11 must also hold for static loading (R = F) and therefore, the sum
of the coefficients of R and F in these equations must be equal to 1/2.

6.4. Method to Compute Shear Demand using a Linear SDOF System

6.4.1. General
Biggs (1964) assumed that the distribution of inertia forces is proportional to the
static deflected shape of a beam (section 6.3). But experiments and numerical
analysis (chapter 3) show dramatic differences between the initial deformed
shapes of beams loaded rapidly and the deformed shapes of beams loaded
statically (Figure 6.3). For blast load, the initial deformations concentrate near the
supports and the middle portion of the beam remains nearly flat (Figure 6.3b-c).
Next, the consequences of this change in the deformed shape are discussed.

169
6.4.2. Method
Consider a fixed-fixed beam in the linear range, with rapidly applied uniform load
(q) and uniform mass (Figure 6.4a). An idealization of the initial deflected shape
of the beam is shown in Figure 6.4b. Consider the dynamic equilibrium of the left
half of the beam (Figure 6.4c). By symmetry, the shear force at mid-span (S) is
zero. Taking moments about the location of the resultant of inertia forces:

V x(t ) ( M m + M end )

1
F x(t ) L / 4 = 0
2

Eq. 6.12

V = end shear (support reaction)


x = distance from support to the location of the resultant of inertia forces

Mm = mid-span moment
Mend = end moment
F = total applied load
L= beam length (clear span)
Unlike the derivation in section 6.3, here the distance x is assumed to vary with
time. This is to account for the change in the beam shape from no deformation
at t = 0 to a final shape that resembles the static deflected shape. Because the
deflected shape of the beam is changing with time during the initial phase, the
distribution of inertia forces (and, therefore, the location of their resultant) must
also change with time. Defining the beam resistance R (variable with time) as:

R(t ) =

8( M m (t ) + M end (t ))
L

Mm = mid-span moment
Mend = end moment
L= beam length (clear span)

Eq. 6.13

170
Substituting Eq. 6.13 into Eq. 6.12, and solving for V:

V (t ) =

1
L
L
F (t )
R (t ) +

2
8 x(t )
8
x
(
t
)

Eq. 6.14

L= beam length (clear span)


x = distance from support to the location of the resultant of inertia forces

R = beam resistance
F = total applied load
Using Eq. 6.14, the dynamic reaction of a fixed-fixed beam can be calculated for
any deflected shape at a given time t as a function of the location of the resultant
of inertia forces x . It is convenient to express Eq. 6.14 as:
V (t ) = (t ) R(t ) + (t ) F (t )

Eq. 6.15

where:

(t ) =
1

L
8 x(t )

(t ) =
2

8 x(t )

Eq. 6.16a
Eq. 6.16b

To estimate shear demand using Eq. 6.15, the variation with time of the factors
and , and the beam resistance R during the initial phase must be determined.
The assumed variation of the applied load F with time is known (blast pressure is
assumed to decrease linearly with time from p0 at t = 0 to 0 at t = td).
The variation of the factors and with time during the initial phase is a
consequence of the dependence of these factors on the deflected shape of the
beam (Eq. 6.16a-b), which changes over time (in the initial phase). At t = 0 the

171
shape of the beam is completely flat and the location of the resultant of inertia
forces is x = L / 4 . Therefore, = 0.5 and = 0 . If x = L / 3 as in the case of a
triangular deformed shape (plastic range), = 3 / 8 and = 1 / 8 (Biggs, 1964).
For the deformed shape shown in Figure 6.5, x = 71L / 240 , = 0.42, and

= 0.08 . Compare this with = 0.14 for a linear beam with fixed ends and a
deflected shape that resembles the static shape (Eq. 6.11). The range of
variation of is relatively wide. For a beam under blast load, the initial deformed
shape is closer to that shown in Figure 6.3b-c than to the static deflected shape
(Figure 6.3a). The variation of (and , using = 1/2 - ) with time is studied
numerically in section 6.4.3.
The beam resistance R depends on the distribution of internal resisting moments
(Eq. 6.13). The magnitude of these moments varies throughout the dynamic
response of the beam and, therefore, R varies with time. If an equivalent SDOF
oscillator is used to represent the beam, R is equal to the force in the spring at
any time t. In the linear range:
R (t ) = k e (t ) y (t )

Eq. 6.17

ke = spring constant for equivalent SDOF system (beam stiffness)


y = displacement of equivalent SDOF model (beam mid-span deflection)
Beam deflections (y) required to compute R can be obtained using linear SDOF
time-history analysis. Beam stiffness (ke) is assumed to vary with time in Eq.
6.17. This assumption is made to account for the change in the deflected shape
in the initial phase (ke depends on the type of load and deformation of the beam).
The variation of beam stiffness with time is studied in section 6.4.4.

172
6.4.3. Variation of Factor with Time
Before any deformation takes place ( t = 0 ), the shape of a beam is completely
flat and therefore, the location of the resultant of inertia forces is x = L / 4 .
Substituting this value into Eq. 6.16b leads to = 0 at time zero. At a time t
approaching T/4 (where T is the vibration period of the beam), the values of
can be obtained based on the static deflected shape of the element (section 6.3).
According to Biggs (1964), for uniformly distributed mass and load, ranges
between 0.11-0.12 for a simply supported beam, and 0.11-0.14 for a beam with
fixed ends, depending on whether the response of the beam is linear or nonlinear. For simplicity, it can be said that increases from 0 to approximately 1/8
(average upper limit, independent of boundary conditions) as the response of the
beam progresses from no deformation at t = 0 to a shape resembling the static
deflected shape. The time required for a beam to reach its static deformed shape
is believed to depend on its fundamental period of vibration. Here the variation of
with time will be studied as a function of the dimensionless parameter t/T.
To study the variation of with time in the initial phase, finite element (FE)
simulations of linear response of actual and hypothetical fixed-fixed beams to
rapidly applied uniform loads were conducted using LS-DYNA (LSTC, 2005).
Sixteen numerical models of beams with geometric and material properties as
listed in Table 6.1 were developed. This table also lists the vibration periods (T)
of the beams analyzed, ranging between 0.2-20 ms. The types of elements and
boundary conditions used in the models are the same as described in section
3.3.2. Triangular loads (Figure 6.1b) with duration td ranging between 0.01T-0.2T
and peak pressure p0 in the range 5-66 ksi (Table 6.1) were used.
Only linear response was considered. The justification for the use of linear
analysis to study the response of beams to blast load during the initial phase was
discussed in section 6.2. The beam models listed in Table 6.1 were built using
LS-DYNA Material Type 1. This material is included in LS-DYNA to represent a

173
linear isotropic elastic material (LSTC, 2003). The following material properties
were common to all the models:
Poissons ratio:

= 0.33

Mass density:

= 2.52 x 10-4 lbf-s2/in4 (2,700 kg/m3)

For each simulation, the maximum shear force and stress at the end of the beam
in the initial phase (Vmax and max, respectively), and the time at which they occur
(tv, with 0 tv 0.1T) were obtained (Table 6.1). Deflected shapes of the beams
at time tv were also obtained and are shown in Figure 6.6 to Figure 6.10. From
the deflected shapes, the location of the resultants of inertia forces and the factor
(at t = tv) for each beam/load combination analyzed were calculated (Table
6.1). Using the results from the sixteen models analyzed, the variation of with
the ratio tv/T was obtained (Figure 6.11a). A similar plot can be generated by
looking at the variation of the deflected shape and with time (in the initial
phase) in a single beam (Figure 6.11a). For simplicity when computing shear
demand in the initial phase, the variation of with time can be represented with
the bilinear envelope shown in Figure 6.11b. As the duration of the load
approaches zero the factor tends to zero and, therefore, the envelope for
estimation of initial shear demand (Figure 6.11b) becomes more conservative.

6.4.4. Variation of Beam Stiffness with Time


Beam stiffness (ke) in the initial phase is defined using the deflected shape and
the reference load w shown in Figure 6.12a. The beam is idealized as having two
deformable segments of length L ( AB and CD ) connected by a rigid middle
segment ( BC ) of length L' = L 2 L' (Figure 6.12b). The maximum deflection is
assumed constant between B and C. To determine ke, consider the deformation
of AB . The total deflection at B has two components (Figure 6.12c):

174

= 1 + 2

Eq. 6.18

1 = deflection due to shear force at B


2 = deflection due to reference load w acting on AB
Assuming linear behavior:

wL' 4

1 =

24 EI

(downward)

Eq. 6.19

L = length of AB
= constant (L= length of BC )
w = reference uniform load
E = Youngs modulus of beam material
I = Moment of inertia of beam section
As shown in Figure 6.12d, the deflection 2 can be obtained as:

2 = 21 + 22

Eq. 6.20

21 = deflection of a cantilever of length L caused by the reference load w


22 = deflection of a cantilever of length L caused by the end moment M
If the slope at B is zero (as assumed in Figure 6.12b), and for a linear beam:

M =

wL' 2
6

21

wL' 4
=
8 EI

22 =

(counter-clockwise)

(downward)

wL' 4
(upward)
12 EI

Eq. 6.21
Eq. 6.22
Eq. 6.23

175
Where:
w = reference uniform load
L = length of AB
E = Youngs modulus of beam material
I = Moment of inertia of beam section
Substituting Eq. 6.22 and Eq. 6.23 into Eq. 6.20:
2 =

wL' 4
(downward),
24 EI

Eq. 6.24

and substituting Eq. 6.19 and Eq. 6.24 into Eq. 6.18:
=

wL' 4
(1 + )
24 EI

Eq. 6.25

= constant (L= length of BC , with L= length of AB )


From Figure 6.12b:

L ' = L 2 L '

Eq. 6.26

= L' / L

Eq. 6.27

Eq. 6.28

Define:

From Eq. 6.25-6.26:

Combining Eq. 6.25 and Eq. 6.27-6.28:

wL4 4 1
1
24 EI

Eq. 6.29

176
Where:
w = reference uniform load
L = beam length (clear span)
E = Youngs modulus of beam material
I = Moment of inertia of beam section
= constant = L/L, where L is the length of AB (Figure 6.12b)
From Eq. 6.29, the stiffness of the beam (spring constant) based on the deflected
shape shown in Figure 6.12b is:

ke =

24 EI
1

3
(1 ) 3
L

Eq. 6.30

Eq. 6.30 can be used to estimate the stiffness of the beam as the deflected
shape varies over time. The difficulty in using Eq. 6.30 is related to the fact that
the variation of with time is unknown. At an infinitesimal time t approaching
zero, little deformation has taken place and the beam has a nearly flat shape. In
this case 0 and the beam stiffness ke tends to infinity. On the other hand,
once the beam has reached a deformed shape similar to the static deflected
shape (for instance, for t > T/4), = 0.5 (the total length of the beam has bent)
and the expression obtained from Eq. 6.30 for ke matches the expression for the
stiffness of a fixed-fixed linear beam subjected to uniform load (ke = 384EI/L3).
Thus, during the initial phase increases from 0 to 0.5 and the beam stiffness
decreases from infinity to the stiffness determined based on the static deflected
shape. Define a stiffness increase factor ():

ke
(k e ) static

24 EI
1

3
1
(1 ) 3
L
=
=
384 EI
16(1 ) 3
3
L

Eq. 6.31

177
Where:
L = beam length (clear span)
E = Youngs modulus of beam material
I = Moment of inertia of beam section
= constant = L/L, where L is the length of AB (Figure 6.12b)
The beam stiffness during the initial phase can be expressed in terms of the
static stiffness (ke)static. For a fixed-fixed beam, from Eq. 6.31:

k e = (k e ) static =

(k e ) static
16(1 ) 3

Eq. 6.32

For instance, if = 0.2 , = 9.8 and the initial stiffness is approximately 10 times
higher than the stiffness obtained from the static deflected shape. An upper limit
to the stiffness increase factor must be established through calibration. The
lower limit to is 1. These limits define the range of variation of beam stiffness
within the time interval 0 t 0.1T. This variation is required for conducting stepby-step analyses of SDOF systems to obtain the initial response (deflections and
resistance forces) of beams to blast load (section 6.4.2).
To calibrate the factor (Eq. 6.31), FEA results were used as reference.
Maximum shears (in the interval 0 t 0.1T) at the ends of fixed-fixed beam
models subjected to rapidly applied uniform load were computed using LS-DYNA
(LSTC, 2005) simulation. Results for the numerical models previously developed
to study the variation of with time (section 6.4.3) were also used for calibration
of . The geometric, material, and load properties of these models were
described in section 6.4.3 and summarized in Table 6.1. The most relevant
properties of these models are listed again in Table 6.2 for an easier examination
and comparison of results. Beam response was assumed linear in all the models.
Results for maximum shear demand in the initial phase (Vmax and max) obtained
from the LS-DYNA simulations are listed in Table 6.2.

178
The beams analyzed using LS-DYNA were also modeled and analyzed using
SDOF systems with variable stiffness (ke) during the initial phase (Table 6.2). The
variation of ke with time was selected so that the results for initial shear obtained
using Eq. 6.15 and SDOF analysis were close to the results obtained from FEA.
The best approximation was obtained by assuming that the stiffness increase
factor (Eq. 6.31) decreases linearly from 10 to 1 ( increases from 0.2 to 0.5) as
time progresses from 0 to 0.1T (Figure 6.12e). Equivalently, ke (Eq. 6.32)
decreases linearly from 3840EI/L3 at t = 0 to 384EI/L3 at t = 0.1T (fixed-fixed
beam). The masses of the equivalent SDOF systems (Me) were computed using
the factors proposed by Biggs (1964). For a fixed-fixed beam, Me is taken as 0.77
times the total mass of the beam. The assumption of constant Me in the initial
phase was made for simplicity, although it is not physically consistent with the
previously mentioned change in the deflected shape of the beam.
The -method proposed by Newmark (1959) was used for dynamic analyses of
the equivalent SDOF systems subjected to blast load. The analyses were done
using an algorithm implemented in Mathcad (PTC, 2007) and presented in
Appendix C. Maximum shear demands in the initial phase (Vmax and max)
obtained from calculated spring forces and using Eq. 6.15 are listed in Table 6.2.
The mean of the ratio of calculated demand using Eq. 6.15 (SDOF) to calculated
demand using LS-DYNA is 0.90. Details of the use of the method to estimate
initial shear demand in beams subjected to blast load are given in Appendix C.

6.5. Computation of Shear Demand using Timoshenko Beam Theory

6.5.1. General
Linear dynamic behavior of beams is typically described using either BernoulliEuler or Timoshenko beam theory. In the Bernoulli-Euler theory it is assumed
that plane sections remain plane after deformation, and shear deformations and

179
rotational inertia are not considered. The beam is assumed to be infinitely stiff in
shear. The Timoshenko theory includes the effects of shear deformation and
rotational inertia. The beam is assumed to have a finite stiffness in shear.
It has been shown that the Bernoulli-Euler theory works well only for the lower
modes of vibration (Prescott, 1942). When higher modes need to be considered,
as in the case of impulsive load (blast or impact), the Bernoulli-Euler theory does
not yield good results. This theory leads to the result that a wave of infinitesimal
wave-length propagates through a beam with infinite velocity (Volterra and
Zachmanoglou, 1965), which is not physically possible. This implies that shear
forces at the ends of a beam loaded impulsively tend to infinity as the load
duration (td) decreases. Therefore, the Bernoulli-Euler theory is not adequate to
estimate shear demand in beams subjected to loads of short duration (td < 0.1T).
The beam theory proposed by Timoshenko (1937) includes the effects of shear
deformations and rotational inertia, which are more relevant in the higher modes.
When this theory is used, a shear wave of infinitesimal wave-length has a
bounded propagation velocity, and calculated shear forces at the ends of a beam
loaded impulsively remain finite as the duration of the load approaches zero
(Graff, 1991). This fact supports the use of the Timoshenko theory to investigate
initial shear demand in beams subjected to blast load.

6.5.2. Differential Equations of Motion


Consider the beam shown in Figure 6.13a. To include the effect of shear
deformation in the dynamic response of the beam, the total rotation (slope) at a
given section,

y
, is expressed as (Figure 6.13b):
x
y
= +
x

Eq. 6.33

180
Where:
y = beam deflection at a section located at a distance x from a support
= beam rotation caused by bending
= beam rotation caused by shear
The equations of motion of the beam are obtained considering the dynamic
equilibrium of segment AB in Figure 6.13c. These equations are available in the
literature (e.g., Timoshenko, 1937; Achenbach, 1973; Graff, 1991). Following the
notation and format used by Graff (1991), the governing equations of the beam
can be written as:

Translational motion:

2 y
2 y
2 + A 2 = q ( x, t )
GA '
t
x x

Eq. 6.34a

Rotational motion:

2
2

y
GA ' + EI 2 = I 2
x
t

Eq. 6.34b

A = beam cross-sectional area


I = moment of inertia of beam cross section
E = Youngs modulus
G = shear modulus
= mass density
= shear coefficient (depends on the cross section shape)
x = position along beam axis (distance from beam support)
t = time at which beam response is evaluated
q = q(x,t) = applied distributed lateral load
y = y(x,t) = total lateral deflection of beam
= (x,t) = beam rotation due to bending

= ( x, t ) =

y
= beam rotation due to shear
x

181
6.5.3. Discrete Approximation and Numerical Solution of the Equations of Motion
Using Finite Differences
The dynamic response in the linear range of a beam loaded impulsively can be
estimated using Eq. 6.34a-b. These two hyperbolic partial differential equations
(PDE) are coupled in y and (a different selection of variables for the solution of
Eq. 6.34 is possible), and must be solved simultaneously. These equations are
solved here numerically using an explicit finite-difference method. A rectangular
grid is used for discrete approximation of the continuous two-dimensional spacetime domain of Eq. 6.34 (Figure 6.14). The spatial dimension corresponds to the
position along the axis of the beam (x), and the temporal dimension corresponds
to the time at which dynamic response is evaluated (t). Constant intervals h = x
and k = t are used to discretize the ranges of variation of x (0 x L/2, where L
is the beam length; only one half of the beam is analyzed because of symmetry)
and t (0 t ta, where ta is the analysis time), respectively. A total of n+1
intervals are used in the x-direction in the grid shown in Figure 6.14, and p+1
intervals are used in the t-direction. The choice of number of intervals in the x
and t directions is made for convenience in the implementation of the numerical
solution (Appendix D). In the continuous domain, the mathematically exact
solutions of Eq. 6.34 are y(x,t) and (x,t). At point (xi,tj) in the grid used for
discrete approximation, the solutions are:

y ( xi , t j ) = y (ih, jk ) = y i , j

Eq. 6.35a

( xi , t j ) = (ih, jk ) = i , j

Eq. 6.35b

where i = 0,1,2,,n+1 and j = 0,1,2,,p+1. Numerical approximations of y(x,t)


and (x,t) are represented by Y(x,t) and (x,t). Discretely, at point (xi,tj):

Y ( xi , t j ) = Y (ih, jk ) = Yi , j

Eq. 6.36a

( xi , t j ) = (ih, jk ) = i , j

Eq. 6.36b

182
The derivatives in Eq. 6.34 can be approximated using finite differences. Using
second-order central difference approximations (Ames, 1969):

y
x

x
y
t

t
2 y
x 2
2
x 2
2 y
t 2

2
t 2

i, j

i, j

i, j

i, j

i, j

i, j

i, j

i, j

y i +1, j y i 1, j

Eq. 6.37a

2h

i +1, j i 1, j

Eq. 6.37b

2h
y i , j +1 y i , j 1

Eq. 6.38a

2k

i , j +1 i , j 1

Eq. 6.38b

2k

y i +1, j 2 y i , j + y i 1, j

Eq. 6.39a

h2

i +1, j 2 i , j + i 1, j
h

Eq. 6.39b

y i , j +1 2 y i , j + y i , j 1

Eq. 6.40a

k2

i , j +1 2 i , j + i , j 1

Eq. 6.40b

k2

Substituting Eq. 6.35 and Eq. 6.37-6.40 into Eq. 6.34, discrete expressions for
the equations of motion of the beam are obtained:
1
1

GA ' ( i +1, j i 1, j ) 2 ( y i +1, j 2 y i , j + y i 1, j )


h
2h

A
+ 2 ( y i , j +1 2 y i , j + y i , j 1 ) = qi , j
k

1
EI
GA ' ( y i +1, j y i 1, j ) i , j + 2 ( i +1, j 2 i , j + i 1, j )
2h
h
I
= 2 ( i , j +1 2 i , j + i , j 1 )
k

Eq. 6.41a

Eq. 6.41b

183
Where qi,j is the discrete representation of the distributed load acting on the
beam, q(x,t). Eq. 6.41a governs translational motion and Eq. 6.41b governs
rotational motion of the beam. Using the discrete approximations Yi,j and i,j (Eq.
6.36) instead of the discrete representations of the exact solutions yi,j and i,j in
Eq. 6.41, explicit formulas for the approximate beam deflection (Y) and bending
angle () at different points in the beam and at different times are obtained. For
a point located at a distance xi = ih from the support, the solution (Y and ) at a
time tj+1 = (j+1)k is calculated from the solution at previous times at that point and
the adjacent points (Figure 6.14). The recurrence formulas obtained are:

Yi , j +1 =

i , j +1 =

G ' m 2
(Yi +1, j + Yi 1, j ) + 21

k2
kG ' m
qi , j
(i +1, j i 1, j )
+
2
A

G ' m 2

Yi , j Yi , j 1

k 2 GA ' 2 Em 2
(i +1, j + i 1, j ) + 2

kGA ' m
+
(Yi +1, j Yi 1, j )
2 I
Em 2

i , j i , j 1

Eq. 6.42a

Eq. 6.42b

where m is the mesh ratio and is defined as:


m = k /h

Eq. 6.43

k = interval of discretization of the grid (mesh) in the t-direction


h = interval of discretization of the grid (mesh) in the x-direction
It can be shown that if m 1 , Eq. 6.42a-b converge (Ames, 1969). If the blast
load applied to the beam is assumed to be uniform, the load intensity q at a point
in the beam does not depend on the position of the point (x) but only on time, i.e.,
q = q(t). The discrete representation of the load becomes just qj and is obtained
from the blast-pressure-vs.-time curve, assumed to be known (Figure 6.1b).

184
6.5.4. Initial and Boundary Conditions
Initial and boundary conditions of the problem must be specified before Eq. 6.42
can be used to obtain the approximate solution (Y and ). The initial conditions
(IC) of a beam loaded dynamically depend on the deformed shape and velocity
of the beam at t = 0 (time at which analysis of beam response starts). The
boundary conditions (BC) depend on how the ends of the beam are supported.
The focus of this chapter is shear demand in a fixed-fixed beam initially at rest
(and undeformed), subjected to uniform blast load. In this case, deflections,
rotations, and velocities are zero at every point in the beam at t = 0. The IC of the
beam can be written in continuous and discrete notation as:
yi ,0 = 0

Eq. 6.44a

( x,0) = 0

i ,0 = 0

Eq. 6.44b

y
( x,0) = 0
t

y
t

Eq. 6.45a

( x,0) = 0
t

y ( x,0) = 0

=0
i ,0

=0
i ,0

Eq. 6.45b

where i = 0,1,2,,n+1. Approximating the time derivatives in Eq. 6.45 with the
formulas in Eq. 6.38 (forcing equality), and using Yi,j and i,j instead of yi,j and i,j
in Eq. 6.44-6.45, discrete approximate expressions for the IC are obtained:

Yi ,1 Yi , 1
2k
i ,1 i , 1
2k

Yi , 0 = 0

Eq. 6.46a

i , 0 = 0

Eq. 6.46b

= 0 Yi , 1 = Yi ,1
= 0 i , 1 = i ,1

Eq. 6.47a
Eq. 6.47b

185
In Eq. 6.47 a false time limit (t = -k < 0) was used. Despite the lack of physical
meaning of Yi,-1 and i,-1, Eq. 6.47a-b are required for computation of the solution
at t = k (j+1 = 1 in Eq. 6.42). Because of the explicit nature of the method used,
solutions at t > k depend on the solution at t = k. Because the beam has fixed
ends, deflections and rotations are zero at x = 0 and x = L, where L is the clear
span. For a prismatic beam with uniform load and material properties, the
deformed shape is symmetric about mid-span. Thus, beam points equidistant
from the mid-span section but on opposite sides of the beam have identical
deflections, and rotations of the same magnitude but opposite direction. Beam
rotations (total slope and bending angle) at mid-span are zero. Taking advantage
of symmetry, only one half of the beam is analyzed here (0 x L/2, with L/2 =
(n+1)*h). The BC of the one-half of the beam analyzed can be written in
continuous and discrete notation as:
y (0, t ) = 0

(0, t ) = ( L / 2, t ) = 0
y
y
(0, t ) = ( L / 2, t ) = 0
x
x

Eq. 6.48

y 0, j = 0

0, j = n +1, j = 0

y
x

=
0, j

y
x

=0
n +1 , j

Eq. 6.49

Eq. 6.50

where j = 0,1,2,,p+1. Approximating the derivative with respect to x in Eq. 6.50


with the formula in Eq. 6.37a (forcing equality), using Yi,j and i,j instead of yi,j
and i,j in Eq. 6.48-6.50, and from the symmetry of the deflected shape, discrete
approximate expressions for the required BC are obtained:
Y0, j = 0

Eq. 6.51

0, j = n +1, j = 0

Eq. 6.52

Yn + 2, j Yn , j
2h

= 0 Yn , j = Yn + 2, j

n , j = n + 2, j

Eq. 6.53
Eq. 6.54

186
6.5.5. Implementation and Application of the Numerical Method
The IC (Eq. 6.46-6.47) and BC (Eq. 6.51-6.54) of the problem, along with the
recurrence formulas given in Eq. 6.42 must be implemented in a computer code
and solved to obtain approximations to the beam deflection (y) and bending
angle (). Once the solution has been obtained in all the points of the grid shown
in Figure 6.14, the bending moment (M) and shear force (V) at a given section of
the beam and at a given time can be calculated using the expressions obtained
from linear beam theory:
M = EI

y
V = ' AG = ' AG

Eq. 6.55a
Eq. 6.55b

A = beam cross-sectional area


I = moment of inertia of beam cross section
E = Youngs modulus
G = shear modulus
= shear coefficient (depends on the cross section shape)
x = position along beam axis
y = total lateral deflection of beam
= beam rotation due to bending

y
= beam rotation due to shear
x

Solving Eq. 6.55 in a discrete fashion (using Yi,j and i,j instead of yi,j and i,j,),
and approximating the derivatives with respect to x with the formulas in Eq. 6.37,
V and M can be calculated at all the points of the grid (Figure 6.14). The
numerical method described to obtain deflections, moments, and shears of a
linear beam using Timoshenko theory was coded and executed using Matlab

187
(The Mathworks, 2007), and was used for evaluation of initial response of beams
to blast load. The Matlab code used is presented in Appendix D.

6.5.6. Results for Initial Shear Demand in Beams Subjected to Blast Load
The linear models of beams subjected to blast load listed in Table 6.2 (studied in
section 6.4 using SDOF analysis and LS-DYNA simulation) were analyzed
numerically using Timoshenko beam theory (TBT). Computations were done
using the algorithm presented in Appendix D. Maximum shear demands in the
initial phase (Vmax and max) obtained from TBT are listed and compared with FEA
(LS-DYNA) results in Table 6.2. The mean of the ratio of calculated demand
using TBT to calculated demand using LS-DYNA is 1.0. Typical results for the
deflected shapes of the beams at the times of maximum shear in the initial phase
(tV) obtained from TBT are compared with LS-DYNA results in Figure 6.15. The
results from both methods are in good agreement.

6.5.7. Comparison of Results from the Proposed Numerical Method with Results
from an Available Analytical Method
Motivated by experiments done by Slawson (1984), Ross and Krawinkler (1985)
developed an analytical method to estimate initial moment (M) and shear (V)
demand in beams subjected to impulsive load. In their method the equations of
motion proposed by Timoshenko were solved using the assumption that the
solution is characterized by harmonic motion. Thus, beam motion can be
described using superposition of orthogonal normal modes of vibration.
Because there are two dependent variables in the Timoshenko formulation
(beam rotation related to bending and beam rotation related to shear), two types
of normal modes were considered. Estimates of moment demand typically
require the use of at least the first seven normal modes, while estimates of shear
demand require the use of at least the first 21 normal modes.

188
Typical results obtained with the method proposed by Ross and Krawinkler are
shown in Figure 6.16. These results were obtained for a one-way RC slab with
fixed ends. The depth and clear span of the slab were assumed to be 7.25 in.
and 44.75 in., respectively. The moment strength of the slab (Mc) was assumed
to be 32.1 kip-in. and its shear strength (Vc) was assumed to be 13.5 kip. The
applied load was idealized as shown in Figure 6.1b with a peak pressure (p0)
approximately equal to 5ksi (35 MPa) or 2 ksi (14 MPa). The dwell time (td) was
fixed at 1 ms and the rise time (tr) was assumed to be zero. The results shown in
Figure 6.16 were normalized with respect to Mc and Vc. The times at which V/Vc
and M/Mc are equal to 1 (t and t, respectively) are indicated in Figure 6.16. The
slab was assumed to fail in shear if t < t. Thus, shear failure is expected only in
the slab subjected to p0 = 35 MPa. Figure 6.16 suggests that at early times,
shear increases at a higher rate than moment, and at later times, moment
increases faster than shear.
Similar variations of V/Vc and M/Mc with time (Figure 6.17), and similar
observations about the initial response of the slab to blast load were obtained
using the numerical method described in sections 6.5.3-6.5.5. Computations
were done using the algorithm presented in Appendix D. Comparing Figure 6.16
and Figure 6.17 shows that the numerical method described in sections 6.5.36.5.5 and the analytical method proposed by Ross and Krawinkler (1985)
produce similar results.

6.6. Application of the Methods Proposed to Estimate Shear Demand to the


Analysis of Initial Response of RC Slabs Tested under Blast Load

6.6.1. Description of Tests


Eleven quarter-scale RC box-type structures (Figure 6.18) were tested under
blast loads at the US Army Engineer Waterways Experiment Station, WES

189
(Slawson, 1984). The box elements had steel-reinforced walls and floor/roof
slabs. The floor/roof slabs were flat. The structures were buried and tested in a
sand backfill. The purpose of the tests was to investigate the possibility of shear
failure in shallow-buried RC box elements subjected to nuclear blast. Highexplosive primacord in a charge cavity was placed over the structure to
simulate the blast load resulting from a nuclear detonation. Compressive strength
of concrete, structural stiffness, reinforcement ratio, and charge density were the
parameters varied in the tests.
Tests were conducted in two series: FY81 and FY82. The main differences
between the two series were compressive strength of concrete (fc) and stiffness
of the box element (or aspect ratio of the walls/slabs of the box element). In
series FY81 fc was in the range 4-6 ksi and in series FY82 fc was in the range 78 ksi. The lower compressive strength of concrete in series FY81 led to
premature failure of the box elements. According to Slawson (1984), this failure
included considerable concrete crushing at the support (of the roof slab) which
allowed reinforcement steel pullout. A typical posttest view of a test element in
series FY81 is shown in Figure 6.19. Anchorage failure occurred before the roof
slab reached its full flexural capacity in the tests in series FY81. Therefore, only
the tests in series FY82 are considered here for analysis. Dimensions and
material properties of the box elements in series FY82 are listed in Table 6.3.

6.6.2. Load and Structure Idealizations for Analysis


The estimation of pressures caused by the explosive charge on the structure is
not simple. The design and configuration of the charge cavity were complex.
The structure was buried and surrounded by soil. Attenuation of blast pressures
and wave reflections may have occurred. Slawson (1984) did analyses to obtain
the peak applied pressures (p0) and the durations (td) of the blast loads on the
roof slabs of the structures from data recorded in the tests. His results were used

190
here to idealize the blast load on the roof slab as shown in Figure 6.1b. The
parameters p0 and td are listed in Table 6.3 for each test.
The most important elements of the box-shaped test specimens were the roof
slabs. These elements were idealized as isolated one-way slabs subjected to
uniform loads and with fixed ends (Figure 6.20).

6.6.3. Estimated Period and Shear Strength of Roof Slabs


Following the work by Biggs (1964), Slawson (1984) calculated the periods of
vibration of the roof slabs to be between 2.4 and 2.7 ms (Table 6.4). The
effective spring constant suggested by Biggs for a fixed-fixed beam under
uniform load was used as the stiffness of the roof slab for uniform load. The
moment of inertia of the roof slab was taken as the average of the moments of
inertia of the uncracked and cracked sections.
Following the work by Karagozian and Case (1973), and Murtha and Crawford
(1981), and using his own test results, Slawson (1984) estimated unit dynamic
shear strengths (vnd) of the roof slabs that he referred to as direct-shear
strengths. Estimates of vnd in all tests are listed in Table 6.4.

6.6.4. Observed Behavior and Failure Mechanisms


A brief description of the most relevant damage observed in each test is listed in
Table 6.5. Deformation and/or failure modes of the roof slabs were obtained
using high-speed video and from posttest photographs. The observed
deformation/damage in the slabs was dominated by shear. The slabs either had
concentration of deformations near the supports (vertical walls of the box
elements) as shown in Figure 6.21a, or were completely severed at the supports
as shown in Figure 6.21b. Slabs that were severed at the supports were
subjected to blast loads larger than those applied to other slabs.

191
6.6.5. Initial Deformation and Shear Demand
The initial response of the roof slabs to blast load in series FY82 (Table 6.3) was
analyzed numerically using the methods presented in sections 6.4 (linear SDOF
system) and 6.5 (Timoshenko beam theory, TBT). Computations were done
using the algorithms presented in Appendix C (linear SDOF system) and
Appendix D (TBT). The two methods used assume elements to have linear
response. The stiffness of slabs was based on a moment of inertia equal to the
average of the moments of inertia of the uncracked and cracked sections. This
was handled numerically by applying a 50% reduction to the modulus of elasticity
of the concrete, Ec (Table 6.4), calculated as 57,000 f ' c [psi].
Maximum initial shear demands in the roof slabs (max) obtained using TBT and
analysis of a linear SDOF system are listed in Table 6.5 for all tests and plotted
versus test number in Figure 6.22. The results from these two methods are
consistent. The mean of the ratio of max calculated using an SDOF system
(section 6.4) to max calculated using TBT (section 6.5) is 0.9. Deflected shapes
at times of maximum initial shear resisted by the slabs calculated using TBT are
shown in Figure 6.23 and Figure 6.24.

6.6.6. Initial Shear Demand vs. Observed Behavior and Estimated Dynamic
Shear Strength
Figure 6.22 shows calculated values of max in each test. The horizontal line in
the figure separates results from specimens that were severed at the supports
from other results. This graph suggests that the dynamic shear strength of the
roof slabs was approximately 7 ksi. This inference is consistent with the results
reported by Slawson (1984) as direct-shear strengths (vnd, Table 6.4).
The inferred resistance is an order of magnitude higher than what would be
expected for static shear. No other physical evidence from dynamic tests is

192
available to justify these results. The following discussion offers a view about the
plausibility of the results obtained.
The deflected shapes of the roof slabs at times of maximum initial shear (Figure
6.23 and Figure 6.24) show that the length of the zone where deformations
concentrated ranged from 0.75 to 1.5 times the effective depth. This is based on
linear analysis. Behavior of the slabs in the tests is not expected to have been
linear up to failure and shear deformations may have concentrated in zones even
shorter than those calculated using linear TBT (Figure 6.23 and Figure 6.24).
This observation implies that the aspect ratios of the portions of the roof slabs
providing shear resistance were low. It is well known that the shear strength of
RC elements increases as aspect ratio decreases (Wood, 1990). It seems
plausible that at the limit (i.e., for an aspect ratio approaching zero) the
resistance to shear approaches the resistance provided by the mechanism
referred to as shear friction. Assuming a coefficient of friction of 1.4, the static
shear-friction resistance (vnf) of the roof slabs could be estimated using the
design expression given by ACI-318 (2008) as:
d
v nf = 1.4 ( + ' ) f y
h

[ psi ]

Eq. 6.56

= bottom (positive) reinforcement ratio


= top (negative) reinforcement ratio
d = effective depth of roof slab
h = thickness of roof slab
fy = yield stress of reinforcement
Values of , , d, h, and fy are listed in Table 6.3. The static shear-friction
resistances of the roof slabs calculated using Eq. 6.56 are:
- For specimens DS2-1, DS2-2, DS2-3:

vnf = 1.5 ksi

- For specimens DS2-4, DS2-5, DS2-6:

vnf = 2.0 ksi

193
Birkeland (1966) collected available test data on ultimate shear stresses in
reinforced concrete joints for different reinforcement ratios. Phillips (1972)
reported test results on the strength of concrete construction joints subjected to
pure shear. Phillips presented his results along with the data collected by
Birkeland in the plot shown in Figure 6.25. The results in this figure show that the
design expression proposed by ACI is conservative and may underestimate
shear-friction resistance by approximately 1/3 in average. It follows that the static
shear strength of the roof slabs could range from 2 to 3 ksi. Comparing these
results with the inferred dynamic shear strength of approximately 7ksi, it is
concluded that strain-rate effects must have led to an increase in strength of
approximately 200%. This is larger than the strength increase observed in the
tests reported in chapter 4 and discussed in chapter 5 (60%). The difference in
increase factors may be related to differences in strain rates. Shear failures in the
tests described in chapter 4 were observed to occur approximately 1 to 2 ms
after impact. Based on computations of shear demand reported by Slawson
(1984) and computations of shear demand using TBT (section 6.6.5), maximum
shears (and, therefore, onset of shear failure) in the roof slabs may have taken
place within 0.3 ms after detonation.
The implication of these observations is that shear strength varies depending on:
(a) Strain rate, and
(b) aspect ratio of the elements resisting shear (which is in turn expected to
vary with time as described in section 6.4).
Not enough data are available to estimate the ranges of variation of shear
strength with these parameters. Therefore, given the uncertainties involved in the
estimation of dynamic shear strength, the results presented in this section do not
lead to absolute conclusions. Nevertheless, the proposed methods (sections 6.4
and 6.5) provided estimates of shear demand that are consistent with the
experimental results reported by Slawson (1984).

194

Table 6.1 Numerical Models Used for Analysis of Initial Response of Beams to Blast Load Variation of with Time
Beam ID Dimensions

E
(ksi)

T
(ms)

td
(ms)

p0
(ksi)

td/T

Vmax

max

(LS-DYNA)

(LS-DYNA)

tv
(ms)

tV/T

(kip)

(ksi)

1"x3/16"x8"

10000

1.7

0.02

40

0.012

12.9

68.8

0.01

0.006

0.032

1"x3/16"x4"

10000

0.4

0.02

40

0.048

12.9

68.8

0.01

0.024

0.059

1"x1/4"x8"

10000

1.3

0.02

51

0.016

19.1

76.4

0.01

0.008

0.035

1"x1/4"x4"

10000

0.3

0.02

51

0.064

19.1

76.4

0.01

0.032

0.064

1"x3/8"x8"

10000

0.8

0.02

66

0.024

30.9

82.4

0.01

0.012

0.039

1"x3/8"x4"

10000

0.2

0.02

66

0.095

30.9

82.4

0.01

0.048

0.069

3"x6"x60"

4000

4.6

0.3

0.065

88.5

4.9

0.20

0.043

0.063

3"x6"x60"

4000

4.6

0.3

10

0.065

177

9.8

0.20

0.043

0.063

3"x6"x60"

4000

4.6

0.3

20

0.065

354

19.7

0.20

0.043

0.063

10

3"x6"x60"

4000

4.6

0.15

20

0.033

236

13.1

0.15

0.033

0.053

11

12"x18"x216"

4000

20

0.050

1180

5.5

0.80

0.040

0.065

12

12"x18"x216"

4000

20

36

0.050

8150

37.7

0.60

0.030

0.054

13

12"x18"x216"

4000

20

0.100

1880

8.7

1.20

0.060

0.082

14

12"x18"x216"

4000

20

36

0.100

11530

53.4

1.20

0.060

0.083

15

12"x18"x216"

4000

20

0.150

2393

11.1

1.60

0.080

0.095

16

12"x18"x216"

4000

20

0.200

2771

12.8

2.00

0.100

0.108

194

195

Table 6.2 Numerical Models Used for Analysis of Initial Response of Beams to Blast Load Maximum Shear Demand
Beam
Dimensions
ID

E
(ksi)

td
(ms)

p0
(ksi)

Vmax

max

Vmax

max

(LS-DYNA)

(LS-DYNA)

(SDOF)

(SDOF)

Vmax (SDOF)/
Vmax (LS-DYNA)

(kip)

(ksi)

(kip)

(ksi)

Vmax

max

(TBT)

(kip)

(TBT)

(ksi)

Vmax (TBT)/
Vmax (LS-DYNA)

1"x3/16"x8"

10000

0.02

40

12.9

68.8

8.2

43.7

0.64

12.9

68.8

1.00

1"x3/16"x4"

10000

0.02

40

12.9

68.8

11.3

60.3

0.88

12.9

68.8

1.00

1"x1/4"x8"

10000

0.02

51

19.1

76.4

11.9

47.6

0.62

19.3

77.2

1.01

1"x1/4"x4"

10000

0.02

51

19.1

76.4

19.2

76.8

1.01

19.3

77.2

1.01

1"x3/8"x8"

10000

0.02

66

30.9

82.4

22.0

58.7

0.71

31.1

82.9

1.01

1"x3/8"x4"

10000

0.02

66

30.9

82.4

32.8

87.5

1.06

31.1

82.9

1.01

3"x6"x60"

4000

0.3

88.5

4.9

85

4.7

0.96

85

4.7

0.96

3"x6"x60"

4000

0.3

10

177

9.8

170

9.4

0.96

170

9.5

0.96

3"x6"x60"

4000

0.3

20

354

19.7

340

18.9

0.96

341

18.9

0.96

10

3"x6"x60"

4000

0.15

20

236

13.1

190

10.6

0.81

231

12.8

0.98

11

12"x18"x216" 4000

1180

5.5

1150

5.3

0.97

1330

6.2

1.13

12

12"x18"x216" 4000

36

8150

37.7

6900

31.9

0.85

8090

37.5

0.99

13

12"x18"x216" 4000

1880

8.7

1970

9.1

1.05

1930

8.9

1.03

14

12"x18"x216" 4000

36

11530

53.4

11840

54.8

1.03

11600

53.7

1.01

15

12"x18"x216" 4000

2393

11.1

2320

10.7

0.97

2380

11.0

0.99

16

12"x18"x216" 4000

2771

12.8

2470

11.4

0.89

2750

12.7

0.99

Mean

0.90

1.00

Standard Deviation

0.14

0.04

195

196

Table 6.3 Dimensions and Material Properties of RC Box Elements and Applied Load (Slawson, 1984)
Test

(1)

DS2-1
DS2-2
DS2-3
DS2-4
DS2-5
DS2-6

Inside
Dimension
(clear span)
L
(in)

Wall/
Slab
Thickness
h,
(in)

Wall/Slab
Effective
Depth,
d
(in)

44.8
44.8
44.8
44.8
44.8
44.8

7.2
7.2
7.2
7.2
7.2
7.2

6.4
6.4
6.4
6.4
6.4
6.4

L/d

Compressive
Strength of
Concrete,
fc
(psi)

Yield Stress
of Long.
Reinforc.,
fy
(ksi)

Yield Stress
of Transv.
Reinforc.,
fwy
(ksi)

Long.
Reinforc.
Ratio,
(2)
=

Transv.
Reinforc.
Ratio,
w

7
7
7
7
7
7

6980
7740
7520
7370
7790
7260

79.5
79.5
79.5
67.3
67.3
67.3

66.1
66.1
66.1
66.1
66.1
66.1

0.0075
0.0075
0.0075
0.012
0.012
0.012

0.008
0.008
0.008
0.008
0.008
0.008

Blast Load on Roof Slab


Peak
Duration
Pressure,
(dwell time),
p0
td
(psi)
(ms)

9150
7360
3970
9800
6520
4160

0.66
0.66
0.66
0.66
0.66
0.66

Table 6.4 Periods of Vibration and Dynamic Shear Strengths of Roof Slabs (Slawson, 1984)
Test

Clear Span,
L
(in)

Thickness
h,
(in)

Effective
Depth,
d
(in)

DS2-1
DS2-2
DS2-3
DS2-4
DS2-5
DS2-6

44.8
44.8
44.8
44.8
44.8
44.8

7.2
7.2
7.2
7.2
7.2
7.2

6.4
6.4
6.4
6.4
6.4
6.4

Compressive
Strength of
Concrete,
fc
(psi)

6980
7740
7520
7370
7790
7260

Modulus of
Elasticity of
Concrete,
Ec
(ksi)

4760
5010
4940
4890
5030
4860

Mass
2
(lbf-s /in)

Period of
Vibration,
T
(ms)

0.075
0.075
0.075
0.076
0.076
0.076

2.7
2.6
2.6
2.4
2.4
2.5

(3)

(2)

td/T

*fy
(long.
reinforc.)
(psi)

w*fwy
(transv.
reinforc.)
(psi)

0.25
0.25
0.25
0.27
0.27
0.27

600
600
600
810
810
810

530
530
530
530
530
530

(1)

Only tests in series FY82 (section 6.6.1) are considered for analysis.

(2)

Longitudinal reinforcement was provided in each face (top and bottom) of the walls/slabs of the box element.

(3)

Mass of a 1-in.-wide slab strip (for analysis purposes).

(4)

Estimates of vnd are based on the work by Karagozian and Case (1973), Murtha and Crawford (1981), and Slawson (1984).

Unit Dynamic
Direct-Shear
Strength

vnd(4)
(psi)

6770
7510
7290
7150
7560
7040

196

197

Table 6.5 Estimated Initial Shear Demand vs. Dynamic Shear Strength and Observed Damage in Roof Slabs
(Slawson, 1984)
Test

td/T

Peak Blast
Pressure,
p0
(psi)

Unit Dynamic
Direct-Shear
Strength
(1)

vnd

(psi)

Observed Damage

Maximum Initial
Shear Demand,
vmax (psi)
TBT

SDOF

DS2-1

0.25

9150

6770

Severance at the supports in vertical failure planes.


Pullout and fracture of longitudinal
reinforcement near the supports.

10340

9310

0.90

DS2-2

0.25

7360

7510

Severance at the supports in inclined failure planes.


Fracture of longitudinal and transverse
reinforcement near the supports.

8280

7520

0.91

DS2-3

0.25

3970

7290

Shear deformations near the supports and central portion


relatively flat with 4-1/8 of permanent midspan deflection.

4480

4070

0.91

DS2-4

0.27

9800

7150

Severance at the supports in vertical failure planes.


Pullout and fracture of longitudinal
reinforcement near the supports.

11030

10000

0.91

DS2-5

0.27

6520

7560

Cracking and spalling of concrete over the entire span.


Permanent midspan deflection was approximately 12.

7380

6620

0.90

DS2-6

0.27

4160

7040

Shear deformations near the supports and central portion


relatively flat with 3-1/2 of permanent midspan deflection.

4690

4280

0.91

Mean

0.91
0.01

Standard Deviation
(1)

vmax-SDOF/
vmax-TBT

Estimates of vnd are based on the work by Karagozian and Case (1973), Murtha and Crawford (1981), and Slawson (1984).

197

Pressure

(above ambient air pressure)

198

p0
+
Positive
Phase

td

ta

Negative Phase

Time

Pressure

(above ambient air pressure)

(a) Typical Pressure vs. Time Curve

p0
+
Positive
Phase

ta

td
(b) Idealized Pressure vs. Time Curve

Figure 6.1 Pressure-Time History for Blast Wave

Time

199

q(t)

L
(a) Simply Supported Beam Subjected to Uniformly Distributed Dynamic Load
F(t)=q(t)L

V (t)

V (t)

I (t)

Distribution of
inertia forces

(b) Inertia Forces and Free Body Diagram of the Beam


L/4

F(t)/2
Mm(t)

V (t)

S=0
x=61L/192

I (t)/2

L/2
(c) Dynamic Equilibrium of One Half of the Beam

Figure 6.2 Determination of Dynamic Reactions (Biggs, 1964)

200

Figure 6.3 Static Deflected Shape vs. Initial Deformed Shape under Blast Load Beam with Fixed Ends

201

q(t)

L
(a) Fixed-Fixed Beam Subjected to Rapidly Applied Uniform Load
F(t)=q(t)L
M end (t)

M end (t)
V (t)

V (t)

I (t)
L

Distribution of
inertia forces

(b) Inertia Forces and Free Body Diagram of the Beam


L/4

F(t)/2
M m(t)

M end (t)
V (t)

S=0
x(t)

I (t)/2
L/2

(c) Dynamic Equilibrium of One Half of the Beam

Figure 6.4 Dynamic Reactions in the Initial Phase of Response to Blast Load

202
Undeformed beam

q(t)

3L/5

L/5

L/5

L
Assumed deformed shape

Figure 6.5 Hypothetical Initial Deformed Shape of a Beam under Blast Load

Distance from Support, x (in)


2

0.00
Beam 1

Deflection (in)

0.01

Beam 3
Beam 5

0.02
0.03
0.04
0.05

Figure 6.6 Deflected Shapes of Fixed-Fixed Beams at Time of Maximum Shear


in the Initial Phase Calculated Using LS-DYNA Beams 1, 3, 5

Distance from Support, x (in)


2

0.00

Deflection (in)

0.01

Beam 2
Beam 4
Beam 6

0.02
0.03
0.04
0.05

Figure 6.7 Deflected Shapes of Fixed-Fixed Beams at Time of Maximum Shear


in the Initial Phase Calculated Using LS-DYNA Beams 2, 4, 6

203

Distance from Support, x (in)


15
10

20

25

30

0.00

Deflection (in)

0.05
0.10
0.15
Beam 7

0.20

Beam 8
Beam 9
Beam 10

0.25

Figure 6.8 Deflected Shapes of Fixed-Fixed Beams at Time of Maximum Shear


in the Initial Phase Calculated Using LS-DYNA Beams 7-10

Distance from Support, x (in)


36
54

18

72

90

108

Deflection (in)

1
2
3
Beam 11
Beam 12

Beam 13
Beam 14

Figure 6.9 Deflected Shapes of Fixed-Fixed Beams at Time of Maximum Shear


in the Initial Phase Calculated Using LS-DYNA Beams 11-14

Distance from Support, x (in)


36
54

18

72

90

108

Deflection (in)

1
2
3
4

Beam 15
Beam 16

Figure 6.10 Deflected Shapes of Fixed-Fixed Beams at Time of Maximum Shear


in the Initial Phase Calculated Using LS-DYNA Beams 15-16

204

0.12
0.10

0.08
0.06
0.04

Results from 16 Beam Models at Time of Max. Shear


Results from 1 Beam Model (Beam 7) at Different Times

0.02
0.00
0.00

0.02

0.04

0.06

0.08

0.10

tV/T, t/T

(a) Results from Numerical Models

0.12
0.10

0.08
0.06
0.04

Results from Numerical Beam Models


Bilinear Envelope

0.02
0.00
0.00

0.02

0.04

0.06

0.08

0.10

t/T

(b) Results from Numerical Models and Bilinear Envelope

Figure 6.11 Variation of with Time in the Initial Phase for Numerical Models of
Fixed-Fixed Beams Subjected to Blast Load

205
Reference load, w

Undeformed beam

L'

L'

L
Deflected shape

(a) Assumed Deflected Shape and Reference Load


D

A
O

L'

L'

L
Rigid segment

(b) Idealization of Beam Stiffness


P=wL/2 w

w
A

=
B

L'

1 +

L'

L'

(c) Deformation of Segment AB


w

w
A

L'

M
A
L'

21

A
L'

22

(d) Deformation of Segment AB under Reference Load w

10

1
0.1

t/T

(e) Variation of Stiffness with Time

Figure 6.12 Idealized Beam Deformation and Stiffness in the Initial Phase of
Response to Blast Load

206

L
Beam

q(x,t)

A'

B'

Deflected shape

dx

(a) Beam Subjected to Lateral Dynamic Load

q(x,t)
M

V
M inertia

A'

y
x

CG

(b) Bending and Shear Deformations


at any Section of the Beam

dx

B'

V
f inertia

M
dx
x

V
dx
x

(c) Dynamic Equilibrium of Segment AB

Figure 6.13 Deformations and Dynamic Equilibrium of a Beam According to the


Timoshenko Theory

207

k = t

h = x

207

Figure 6.14 Approximation of the Two-Dimensional Space-Time Domain of the Equations of Motion of a Beam
According to the Timoshenko Theory and Rectangular Grid Used for Numerical Solution Using Finite Differences

208

(a) Beam 13 (Table 5.2)

(b) Beam 14 (Table 5.2)

Figure 6.15 Deflected Shapes of Fixed-Fixed Beams at Time of Maximum Shear


in the Initial Phase Calculated Using TBT and LS-DYNA

209

Figure 6.16 History of Shear and Moment at the Support of Slab Subjected to
Blast Load (after Ross and Krawinkler, 1985)

210

SHEAR

MOMENT

p0 = 35 MPa
tr = 0
td = 1 ms

t t

(a)

MOMENT

p0 = 14 MPa
tr = 0
td = 1 ms

SHEAR

t t

(b)

Figure 6.17 History of Shear and Moment at the Support of Slab Subjected to
Blast Load (computed using TBT and algorithm in Appendix D)

211

Blast load

Roof slab

Wall
48

in.

Floor slab
Principal (longitudinal)
reinforcement

Roof slab

h (typical)

(typical)

Transverse
reinforcement

Temperature
reinforcement

Section A-A
NOTE: Series FY81: L = 48.0 in., h = 5.6 in., d = 4.8 in.
Series FY82: L = 44.8 in., h = 7.2 in., d = 6.4 in.

Figure 6.18 RC Box Elements Tested under Blast Load by Slawson (1984)

212

Figure 6.19 Typical Posttest View of RC Box Element in Series FY81 Concrete
Crushing and Reinforcement Pullout at the Supports of the Roof Slab
(after Slawson, 1984)

Blast pressure, p(t)

h
L
Figure 6.20 Idealization of Roof Slabs of RC Box Elements

213

(a) Concentration of Def ormations near the Supports


(Test DS2-6)

(b) Direct Shear Failure at the Supports


(Test DS2-1)

Figure 6.21 Typical Posttest Views of RC Box Elements in Series FY82


Deformation/Failure Dominated by Shear (after Slawson, 1984)

214

DS2-

DS2-

DS2-

DS2-

DS2-

DS2-

Figure 6.22 Calculated Maximum Initial Shear Demand in Roof Slabs

215

10

15

Distance f rom Support (in)


20
25

30

35

40

45

Def lection (in)

0.00
DS2-1

0.02

DS2-2

DS2-3

0.04
0.06
0.08

Figure 6.23 Deflected Shapes at Times of Maximum Initial Shear Resisted by Roof Slabs Calculated using TBT

10

15

Distance f rom Support (in)


20
25

30

35

40

45

Def lection (in)

0.00
0.02

DS2-4

DS2-5

DS2-6

0.04
0.06
0.08

Figure 6.24 Deflected Shapes at Times of Maximum Initial Shear Resisted by Roof Slabs Calculated using TBT

215

216

Figure 6.25 Shear Strength of RC Joints Subjected to Pure Shear vs.


Reinforcement Ratio (after Birkeland, 1966, and Phillips, 1972)

217

CHAPTER 7. SUMMARY AND CONCLUSIONS

7.1. Summary

7.1.1. Objective and Scope


The objective of the study described in this report is to obtain a simple way to
determine whether a beam that fails in flexure under static load may fail in shear
under blast or fluid impact.
Methods to determine the likelihood of shear failure under dynamic load were
developed from a simple perspective. It was suggested that the critical parameter
is the ratio of shear demand to shear capacity. For dynamic load, estimation of
shear demand requires consideration of inertial forces. This was done in this
study through numerical analysis (chapter 6). The capacity of an element to
resist shear is sensitive to loading rate. The effect of loading rate on shear
capacity was studied experimentally for small reinforced concrete (RC) beams
(chapter 4 and chapter 5).
The scope of this study included (a) construction, instrumentation, and testing of
forty-two small-scale RC beams under static load and impact of a liquid body,
and (b) implementation of numerical methods to compute initial shear demand on
a beam subjected to dynamic load.
This study does not consider axial load acting on a structural element subjected
to blast or impact load. The study is focused on beams with:

218
(1) End rotations fully restrained,
(2) loads perpendicular to the longitudinal axis, and
(3) static shear strengths higher than shears at yield (i.e., beams that can
reach their flexural capacities without failing in shear).
To investigate whether the phenomenon observed by Menkes and Opat (1973)
can also occur for RC beams and for concentrated dynamic loads, experiments
were conducted. The experimental program and its results (CHAPTER 4) are
limited to:
(1) Small-scale RC beams 1-in. deep, with clear spans of 8 in. and 12 in.,
and widths ranging from 2 in. to 6 in.
(2) Beams with continuous top and bottom longitudinal reinforcement
and without transverse reinforcement.
(3) Longitudinal reinforcement ratios () of 0.5% and 1% (top and
bottom), and longitudinal reinforcement yield stresses (fy) of 50 ksi
and 120 ksi.
(4) Compressive strength of concrete (fc) in the range 6-10 ksi.
(5) Ratio of shear span (a) to effective depth (d) equal to 2.7 and 4.
(6) Concentrated load (static and fluid impact) at midspan.
(7) Impact velocity (vp) ranging approximately from 50 m/s to 150 m/s.
(8) Ratio of duration of impact load (td) to calculated period of vibration of
the impacted beam (T) ranging approximately from 0.4 to 1.4.
(9) Ratio of mass of impacting liquid body (m) to effective mass of the
impacted beam (Me) ranging approximately from 0.2 to 0.5.

7.1.2. Hypothesis
Menkes and Opat (1973) conducted experiments on small-scale aluminum
beams subjected to blast load. As the blast load was increased, the failure mode
of the beams changed from flexure to shear. The behavior of the beams in the
tests was studied numerically using finite-element analysis (FEA). Results from

219
FEA (using LS-DYNA) showed that the maximum calculated shear forces were
close to the nominal plastic shear capacities of the beams. The experimental and
numerical evidence pointed towards a simple hypothesis: A beam that fails in
flexure under static load may fail in shear under blast or impact load if the
maximum initial shear demand reaches the capacity of the beam to resist shear.
This view differs from conclusions of previous work which focused on the velocity
imparted by impact or blast as the key parameter defining the mode of failure.

7.1.3. Experimental Investigation of the Response of Small-Scale Reinforced


Concrete (RC) Beams to Fluid Impact
Tests Specimens and Experimental Plan
Experiments were conducted to investigate whether the change in failure mode
observed by Menkes and Opat (1973) in aluminum beams subjected to uniform
blast load can also occur in RC beams subjected to concentrated dynamic load.
The test specimens were small-scale RC beams with longitudinal reinforcement
ratios of 0.5% and 1%, and with no transverse reinforcement. The specimens
had their ends restrained against rotation and were subjected to concentrated
impact load at midspan. The impacting body consisted of a thin-wall aluminum
can filled with water. The ranges of the variables in the experiments (geometric
and material properties of the beams, and impact velocities) are summarized in
section 7.1.1. The test specimens were grouped into eleven series. The
properties of the specimens were identical in a given series. To establish the
mechanisms of failure of the specimens under slowly applied load, static tests
were conducted before the impact tests. A total of forty-two specimens were
tested: eleven under static load and thirty-one under impact load.
Response and Failure Mechanisms
In the static tests, beams from eight series developed flexural mechanisms of
failure and beams from three series failed in shear. Beams that developed

220
flexural mechanisms of failure reached maximum drift ratios exceeding 7%. Their
deformations concentrated near the supports and at midspan. Flexural cracks
formed only at these locations and were initially difficult to observe. As beam
deflections increased, cracks widened (crack widths exceeded 1/16). Based on
the measured load-deflection responses, the full flexural capacities of the beams
were reached. Beams that failed in shear developed wide inclined cracks
(exceeding 1/16 in width) near midspan at drift ratios not exceeding 3% and had
narrow flexural cracks (not exceeding 0.02 in. in width) near the supports.
Specimens with the same nominal properties as those that developed flexural
mechanisms of failure under static load were tested under impact load. The
extent and type of damage in the impact tests changed from one test to the next
depending on beam geometry, reinforcement ratio, reinforcement yield stress,
and impact velocity. All beams exhibited cracking and inelastic deformations. The
state of the specimens after impact ranged from complete disintegration to small
permanent deflections and narrow cracks (not exceeding 0.03 in. in width).
Thirteen specimens had the same deformation/failure mechanisms under impact
and static load (flexural mechanisms). These deformation/failure mechanisms
were observed at low impact velocities (and, therefore, low impact forces).
Eighteen specimens had failure mechanisms under impact load that were
different from the failure mechanisms under static load. These failure
mechanisms were observed at high impact velocities (and, therefore, high
impact forces), and were dominated by inclined cracking and disintegration of the
concrete (initiated by the inclined cracks). The inclined cracks formed at early
stages of response (within 1 ms after impact) and were typically located near the
quarter points and midspan. Cracking patterns (Figure 4.66) and initial damage
(Figure 4.67 to Figure 4.74) in the beams were obtained using a high-speed
camera.

221
7.1.4. Static and Dynamic Shear Strength of Small-Scale RC Beams
Maximum nominal shear stresses in the static tests were lower than 3 f 'c [psi]
(where fc is the compressive strength of concrete, psi) for beams that developed
flexural mechanisms of failure. For beams that failed in shear, the mean static
shear strength was 3.7 f 'c [psi]. This value is consistent with the results by
Chana (1981), who tested statically small-scale RC beams similar to the
specimens in this study. The beams tested by Chana failed in shear at nominal
shear stresses ranging from 3.3 f ' c to 4.6 f ' c [psi].
In the dynamic tests, beams subjected to low-speed impact failed in flexure and
beams subjected to high-speed impact failed in shear. The threshold between
shear and flexural failure was expressed in terms of initial shear stress. Initial
shear stress was computed using the following assumptions:
- The strength and mass of the container were assumed to be negligible.
- The inertia force acting on the small segment of the beam impacted by the
liquid was neglected.
- The relative velocity between the impacting body (with mass m) and the
beam (with effective mass Me) is equal to the impact velocity.
The maximum initial shear force in the beam was taken as F0/2, where F0 is the
maximum initial force exerted on the beam and is estimated using Eq. 5.20:
F0 = w Av p

w = density of impacting liquid (water)


A = cross-sectional area of impacting liquid body
vp = projectile velocity (impact velocity)

Eq. 5.20

222

Maximum initial nominal shear stresses (max0) ranged between 3 f 'c and

14.5 f 'c [psi] in the impact tests. None of the beams with max0 lower than
6 f 'c (11 beams in total) showed inclined cracking and/or disintegration. These
beams developed flexural mechanisms of failure. Eighteen out of twenty beams
with max0 higher than 6 f 'c had inclined cracking and/or disintegration. This
was observed for test specimens that had different dynamic properties and
different ratios of duration of impact load (td) to calculated period of vibration (T).
This evidence suggests that the initial shear failure of beams under impact load
is related to shear strength and does not depend on vibration properties.
The observed threshold applies exclusively to small-scale RC beams with
geometric and material properties similar to those described in section 7.1.1, td/T
ratios ranging from 0.4 to 1.4, and m/Me ratios ranging from 0.2 to 0.5. The ratio
of the inferred dynamic shear strength to the static shear strength of the test
specimens is 1.6.

7.1.5. Response to Impact of Small-Scale RC Beams with Flexural Mechanisms


of Deformation and/or Failure
The dynamic response of specimens that had deformation/failure mechanisms
dominated by flexure was studied using idealized impact loads and equivalent
elasto-plastic SDOF systems. Increases in the resistances of the SDOF systems
because of high strain rates in the reinforcement of the beams were taken into
account. The reinforcement strain rates in the impact tests were estimated to be
in the range 4-30 s-1. Calculated displacement histories of the SDOF systems
were similar to the measured midspan deflection histories of the test specimens.
In general, calculated maximum and permanent midspan deflections were larger
than the measured midspan deflections. The mean ratio of calculated to
measured maximum midspan deflection was 1.11, but calculation errors of 50%

223
were obtained for four specimens. The error was related to the uncertainties
involved in estimating the dynamic resistances of the beams and the force
exerted by the impacting liquid body on a deformable target (RC beam).

7.1.6. Initial Shear Demand in Beams Subjected to Blast Load


The initial phase of response of a beam to blast load was defined as the time
interval 0 t 0.1T, where T is the fundamental period of the beam. Shear
failure in the initial phase must be avoided. Initial shear demand and dynamic
shear strength of the beam must be determined. Two different methods to
estimate initial shear demand on beams subjected to blast load were proposed.
One method was based on the work by Biggs (1964) to compute dynamic
reactions and used linear SDOF analysis. The other method was based on
Timoshenko beam theory and used numerical analysis of dynamic response of a
linear beam. The use of linear analysis in both methods was justified by
experimental evidence that shows that shear failures in elements subjected to
blast load may occur at early times before flexural deformations can be
observed. Brief descriptions of both methods and their results on prediction of
maximum shear demand in numerical beam models are presented below.
Method to Compute Shear Demand using a Linear SDOF System
As suggested by Biggs (1964), the dynamic reaction was assumed to have two
components: one related to the applied load and other related to the resistance
developed by the beam. Resistance forces were calculated from the deflection
response of the beam in the linear range. This response was obtained using an
equivalent linear SDOF system with time-dependent stiffness to account for the
change of the deflected shape of the beam in the initial phase. The variation with
time of the factor giving the contribution of the applied load to the maximum
shear in the beam (support reaction) was established.

224
Using SDOF analysis and Eq. 6.15, maximum shears at the ends of fixed-fixed
hypothetical beams subjected to blast load and in the initial phase were obtained.
Maximum initial shears in the hypothetical beams were also obtained from FEA
(using LS-DYNA). Response was linear in all beam models. The mean of the
ratio of calculated shear demand using Eq. 6.15 (SDOF) to calculated demand
using LS-DYNA was 0.90.
Method to Compute Shear Demand using Timoshenko Beam Theory
The Timoshenko beam theory includes the effects of shear deformations and
rotational inertia on the dynamic response of a beam. These effects are relevant
in the higher modes of vibration, which are excited by impulsive load (blast or
high-speed impact). Thus, the Timoshenko theory was deemed a suitable tool to
investigate initial shear demand in beams subjected to blast load.
The differential equations of motion of a beam according to the Timoshenko
theory are given in Eq. 6.34a-b. Using these equations, the dynamic response in
the linear range of a beam loaded impulsively can be estimated. Eq. 6.34a-b are
two partial differential equations (PDE) coupled in two variables related to
bending and shear deformation, and must be solved simultaneously. These
equations were solved numerically using an explicit finite-difference method. A
rectangular grid was used for discrete approximation of the continuous twodimensional space-time domain of the problem. The spatial dimension was the
position along the axis of the beam (x), and the temporal dimension was the time
at which dynamic response was evaluated (t). Derivatives with respect to x and t
were approximated using second-order central differences.
Explicit formulas for the approximate beam deflection and bending angle at
different points in the beam and at different times were obtained (Eq. 6.42a-b).
These formulas were used along with the initial and boundary conditions (IC and
BC, respectively) of the problem. The IC (Eq. 6.46-6.47) and the BC (Eq. 6.51-

225
6.54) were defined for a fixed-fixed beam initially at rest and undeformed. Thus,
deflections, rotations, and velocities were zero at every point in the beam at zero
time. The numerical method was coded and executed in Matlab. Approximations
to the beam deflection and bending angle were calculated at different points in
the beam and at different times. From this solution, the bending moment and
shear force at different beam sections and times were calculated using the
classical relationships derived in linear beam theory (Eq. 6.55a-b).
The linear models of beams subjected to blast load previously studied using
SDOF analysis and FEA (LS-DYNA simulation) were analyzed numerically using
Timoshenko beam theory (TBT). Maximum initial shears in the beam models
were obtained. The mean of the ratio of calculated shear demand using TBT to
calculated demand using LS-DYNA was 1.0.

7.2. Conclusions
On the basis of the analyses and experimental evidence presented here the
following conclusions are drawn:
1. The change in failure mode (from flexure to shear) at high-intensity
distributed blast load observed by Menkes and Opat (1973) in
aluminum beams was also observed in small-scale RC beams under
concentrated fluid impact load.
2. Previous studies expressed the threshold at which the change in failure
mode occurs in terms of the velocity imparted to the structural element
during the blast or impact. This study shows that the threshold can be
expressed in terms of force instead of velocity. Expressing the
threshold in terms of shear force allows the formulation of a simpler
criterion to distinguish between beams that fail in shear under dynamic
loads and beams that do not. This criterion is as simple as follows:

226
A beam that fails in flexure under static load may fail in shear under
blast or fluid impact load if the maximum initial shear demand reaches
the capacity of the beam to resist shear.
Shear failure may occur before a flexural mechanism is reached,
limiting the energy- dissipation capacity of the element.
3. Small-scale RC beams with longitudinal reinforcement ratios between
0.5% and 1%, and without transverse reinforcement failed in shear
under fluid impact at nominal shear stresses exceeding 6 f 'c [psi],
where fc is the compressive strength of concrete (in psi). This limit is
approximately 60% higher than the static shear strength and does not
depend on the dynamic properties of the beam. Initial shear demand
was taken as one half of the impact force estimated using Eq. 5.20,
which was obtained following the work by Riera (1968).
4. The deformed shape of a beam in the initial phase of response to blast
load (0 t 0.1T, where T is the period of the beam) is dominated by
shear deformations near the supports, with the central portion of the
span remaining nearly flat. This was confirmed by numerical solutions
of the differential equations of motion of a beam subjected to blast load
according to the Timoshenko theory (which includes the effects of
shear deformations and rotational inertia). If no shear failure occurs in
the initial phase, the shape changes as deformations progress from the
supports towards midspan and the beam acquires a shape similar to
the deflected shape under static load at a time longer than 0.1T.
5. Within the ranges described in section 7.1.1, the formulation proposed
by Biggs (1964) to compute support reactions in prismatic beams
loaded dynamically:

227
V (t ) = R(t ) + F (t )

Eq. 7.1

V = maximum shear (support reaction)


R = beam resistance
F = applied load
, = factors that depend on the deflected shape ( + = 1/2)
provided reasonable estimates of shear demand in beams subjected to
blast load after the following modifications were made to account for
the variation of the deformed shape in the initial phase:
The factor was assumed to vary with time as shown in
Figure 6.11 (and was computed as 1/2 ).
The resistance R(t) was obtained from the response of an
equivalent linear SDOF system and was computed as:
R (t ) = k e (t ) y (t )

Eq. 6.17

ke = stiffness of equivalent SDOF system


y = displacement of equivalent SDOF
with:

10k es 1 9 , t 0.1T
k e (t ) =
T

k ,
t 0.1T
es

Eq. 7.2

kes = stiffness of equivalent SDOF system associated


with the static deflected shape of the beam
The values proposed by Biggs for and in Eq. 7.1 are expected to
give an upper bound to the maximum initial shear demand in a beam
subjected to blast load.

228
7.3. Future Work
The discussion of the results from the impact tests (chapter 5) showed that the
dynamic shear strength of the specimens was approximately 60% higher than
the static shear strength. This was obtained for specimens made with small-scale
concrete, with identical depths (1 in.), with no transverse reinforcement, and
without applied axial load. For elements outside the ranges described in section
7.1.1, the dynamic increase factor for shear strength is unknown. There is need
for more experimental data on dynamic shear strength of reinforced concrete
elements. Future experimental programs should consider test specimens with
transverse reinforcement and with different depths. There is also need for data
on response of columns to fluid impact, which means that axial load should be
one of the parameters to study in future investigations. It would be desirable to
conduct fluid impact tests on larger specimens, but this is difficult because of
limitations related to the test setup. Finally, a different material could be used for
the test specimens (e.g., metal beams).
The methods proposed to estimate initial shear demand caused by blast load
(chapter 6) assume the response of the beam to be linear. The proposed
methods may give upper bounds to initial shear demand if non-linear behavior of
the beam occurs in the initial phase. It is suggested that simple alternatives to
account for non-linear response in the initial phase be investigated. The method
that uses an equivalent SDOF system with time-dependent stiffness (section 6.4)
has potential for implementation of non-linear response.

LIST OF REFERENCES

229

LIST OF REFERENCES

Achenbach, J. D. (1973). Wave Propagation in Elastic Solids. American Elsevier,


New York, New York.
ACI-ASCE Committee 326. (1962). Shear and Diagonal Tension, pt. 2. ACI
Journal, 59(2), 277-333.
ACI Committee 318. (2008). Building Code Requirements for Structural Concrete
and Commentary. American Concrete Institute (ACI), Farmington Hills,
Michigan.
Ames, W. F. (1969). Numerical Methods for Partial Differential Equations.
Barnes and Noble, New York, New York.
Arakawa, T. (1969). Allowable Shearing Stress and Shear Reinforcement for RC
Beams. Summaries of Technical Papers of Annual Meeting Architectural
Institute of Japan, 891-892.
Barry, J. E., and Sozen, M. A. (1970). A Note on the Strength Characteristics of
Small-Scale Concrete with Special Aggregates. A Report on a Research
Project Sponsored by the National Science Foundation, Civil Engineering,
University of Illinois, Urbana, Illinois.
Bazant, Z. P., and Zhou, Y. (2002). Why Did the World Trade Center Collapse?Simple Analysis. Journal of Engineering Mechanics, 128(1), 2-6.
Biggs, J. M. (1964). Introduction to Structural Dynamics. McGraw-Hill, New York,
New York.
Birkeland, F. W., and Birkeland, H. W. (1966). Connection in Precast Concrete
Construction. ACI Journal Proceedings, 63(3), 345-368.
Bresler, B., and MacGregor, J. G. (1967). Review of Concrete Beams Failing in
Shear. American Society of Civil Engineers Proceedings, Journal of the
Structural Division, 93(ST1), 343-372.

230
Chana, P. S. (1981). Some Aspects of Modelling the Behaviour of Reinforced
Concrete under Shear Loading. Technical Report 543 - Cement and
Concrete Association.
Clark, E. N. et. al. (1965). Plastic Deformation of Structures, I: Beams. Air Force
Flight Dynamics Laboratory, Technical Report 64-64, Picatinny Arsenal,
Dover, New Jersey.
Cowper, G. R., and Symonds, P. S. (1957). Strain Hardening and Strain Rate
Effects in the Impact Loading of Cantilever Beams. Technical Report No.
28 from Brown University to the Office of Naval Research under Contract
No. 562(10), Providence, Rhode Island.
Davidson, P., and Davis, R. (1975). Mechanical Behavior of 6061-T651
Aluminum. AIAA Journal, 13(12), 1547-1548.
Federal Emergency Management Agency (FEMA). (1996). The Oklahoma City
Bombing: Improving Building Performance through Multi-Hazard
Mitigation. FEMA 277, Washington, D.C.
Federal Emergency Management Agency (FEMA). (2002). World Trade Center
Performance Study: Data Collection, Preliminary Observations, and
Recommendations. FEMA 403, Washington, D.C.
Fu, H. C., Erki, M. A., and Seckin, M. (1991). Review of Effects of Loading Rate
on Reinforced Concrete. Journal of Structural Engineering, ASCE,
117(12), 3660-3679.
Graff, K. F. (1991). Wave Motion in Elastic Solids. Dover, Mineola, New York.
Hanai, N., Umemura, H., and Ichinose, T. (2005). The Factors which Influence
Strength Deterioration of RC Columns Failing in Shear after Flexural
Yielding. Journal of Structural and Construction Engineering, AIJ, 593,
129-136.
Hoge, K. G. (1966). Influence of Strain Rate on Mechanical Properties of 6061T6 Aluminum under Uniaxial and Biaxial States of Stress. Experimental
Mechanics, 6(4), 204-211.
Humphreys, J. S. (1965). Plastic Deformation of Impulsively Loaded Straight
Clamped Beams. ASME -- Transactions -- Journal of Applied Mechanics,
32(1), 7-10.

231
Irfanoglu, A., and Hoffmann, C. M. (2008). Engineering Perspective of the
Collapse of WTC-I. Journal of Performance of Constructed Facilities,
ASCE, 22(1), 62-67.
Jones, N. (1971). Theoretical Study of the Dynamic Plastic Behavior of Beams
and Plates with Finite Deflections. International Journal of Solids and
Structures, 7(8), 1007-1029.
Jones, N. (1976). Plastic Failure of Ductile Beams Loaded Dynamically. Journal
of Engineering for Industry, Transactions of the ASME, 98(B1), 131-136.
Karagozian and Case (K & C). (1973). Construction Joint Test Program. Final
Report to Space and Missile Systems Organization, Norton AFB, Contract
F04701-72-C-0358, Glendale, California.
Karim, M. R., and Hoo Fatt, M. S. (2005). Impact of the Boeing 767 Aircraft into
the World Trade Center. Journal of Engineering Mechanics, ASCE,
131(10), 1066-1072.
Karna, T., Linna, I., Hakola, I., Juntunen, J., and Jarvinen, E. (2005). An apparatus
for Medium Scale Impact Tests. Experimental Vibration Analysis for Civil
Engineering Structures (EVACES) International Conference, Bordeaux,
France.
Keenan, W.A. (1977). Shear Stress in One-Way Slabs Subjected to Blast Load.
Civil Engineering Laboratory, Naval Construction Batallion Center, Port
Hueneme, California.
Kiger, S. A., and Slawson, T. R. (1984). Dynamic Shear Failures in One-Way
Slabs. Proceedings of the Fifth ASCE Engineering Mechanics Division,
Specialty Conference, 1, 671-674.
Krauthammer, T., Bazeos, N., and Holmquist, T. J. (1986). Modified SDOF
Analysis of RC Box-Type Structures. Journal of Structural Engineering,
ASCE, 112(4), 726-744.
Krauthammer, T., Shahriar, S., and Shanaa, H. M. (1990). Response of
Reinforced Concrete Elements to Severe Impulsive Loads. Journal of
Structural Engineering, ASCE, 116(4), 1061-1079.
Li, Q. M., and Jones, N. (1999). Shear and Adiabatic Shear Failures in an
Impulsively Loaded Fully Clamped Beam. International Journal of Impact
Engineering, 22(6), 589-607.

232
Liu, J., and Jones, N. (1987). Experimental Investigation of Clamped Beams
Struck Transversely by a Mass. International Journal of Impact
Engineering, 6(4), 303-335.
Livermore Software Technology Corporation (LSTC). (2003). LS-DYNA Keyword
Users Manual, Version 970, Livermore, California.
Livermore Software Technology Corporation (LSTC). (2005). LS-DYNA v970
r5434a SMP Version, Livermore, California.
Lynn, A. C., Moehle, J. P., Mahin, S. A., and Holmes, W. T. (1996). Seismic
Evaluation of Existing Reinforced Concrete Columns. Earthquake
Spectra, 12(4), 715-739.
Macgregor, J. G., Sozen, M. A., and Siess, C. P. (1965). Strength of
Prestressed Concrete Beams with Web Reinforcement. ACI Journal
Proceedings, 62(12), 1503-1519.
Malvar, L. J. (1998). Review of Static and Dynamic Properties of Steel
Reinforcing Bars. ACI Materials Journal, 95(5), 609-616.
Malvar, L. J., and Ross, C. A. (1998). Review of Strain Rate Effects for Concrete
in Tension. ACI Materials Journal, 95(6), 735-739.
McCann, D. M., and Smith, S. J. (2007). Blast Resistant Design of Reinforced
Concrete Structures. Structure Magazine, April, 22-26.
Menkes, S. B., and Opat, H. J. (1973). Broken Beams. Experimental
Mechanics, 13(11), 480-486.
Mlakar, P. F., Dusenberry, D. O., Harris, J. R., Haynes, G., Phan, L. T., and
Sozen, M. A. (2003). The Pentagon Building Performance Report. ASCE,
SEI, Reston, Virginia.
Moehle, J. P., and Sozen, M. A. (1980). Experiments to Study Earthquake
Response of R/C Structures with Stiffness Interruptions. Civil Engineering
Studies, Structural Research Series No. 482, Civil Engineering, University
of Illinois, Urbana, Illinois.
Murtha, R., and Crawford, J. (1981). Dynamic Shear Failure Predictions of
Shallow-Buried Reinforced Concrete Slabs. Civil Engineering Laboratory,
Naval Construction Batallion Center, Technical Memorandum No. M-5181-04, Port Hueneme, California.

233
Newmark, N. M. (1953). An Engineering Approach to Blast Resistant Design.
Proceedings of the American Society of Civil Engineers (ASCE), 79,
Separate No. 306.
Newmark, N. M. (1959). Method of Computation for Structural Dynamics.
ASCE -- Proceedings -- Journal of the Engineering Mechanics Division,
85(EM3, Part 1), 67-94.
National Institute of Standards and Technology (NIST). (2005). Final Report of
the National Construction Safety Team on the Collapses of the World
Trade Center Towers. NIST NCSTAR 1, Gaithersburg, Maryland.
Omika, Y., Fukuzawa, E., Koshika, N., Morikawa, H., and Fukuda, R. (2005).
Structural Responses of World Trade Center under Aircraft Attacks.
Journal of Structural Engineering, ASCE, 131(1), 6-15.
Parametric Technology Corporation (PTC). (2007). Mathcad v14.0, Needham,
Massachusetts.
Phillips, M. H. (1972). "Horizontal Construction Joints in Cast-in-Situ Concrete.
Master of Engineering Report, Civil Engineering, University of Canterbury,
Christchurch, New Zealand.
Prescott, J. (1942). Elastic Waves and Vibrations of Thin Rods. London,
Edinburgh, Dublin Philosophical Magazine and Journal of Science,
33(225), 703-754.
Pujol, S., Sozen, M. A., and Ramirez, J. A. (2000). Transverse Reinforcement
for Columns of RC Frames to Resist Earthquakes. Journal of Structural
Engineering, ASCE, 126(4), 461-466.
Pujol, S. (2001). FLECHA (Moment-Curvature Response of Reinforced Concrete
Members). Civil Engineering, Purdue University, West Lafayette, Indiana.
Pujol, S., and Brachmann, I. (2007). "Experimental and Analytical Study on the
Response of Barriers to Fluid Impact. Technical Report, Civil Engineering,
Purdue University, West Lafayette, Indiana.
Richart, F. E. (1927). An Investigation of Web Stresses in Reinforced Concrete
Beams. University of Illinois Engineering Experimental Station, Bulletin
No. 166, June, Urbana, Illinois.
Riera, J. D. (1968). On Stress Analysis of Structures Subjected to Aircraft
Impact Forces. Nuclear Engineering and Design, 8(4), 415-426.

234
Ross, T. J., and Krawinkler, H. (1985). Impulsive Direct Shear Failure in RC
Slabs. Journal of Structural Engineering, ASCE, 111(8), 1661-1677.
Shen, W. Q., and Jones, N. (1992). A Failure Criterion for Beams under
Impulsive Loading. International Journal of Impact Engineering, 12(1),
101-121.
Slawson, T. R. (1984). Dynamic Shear Failure of Shallow-Buried Flat-Roofed
Reinforced Concrete Structures Subjected to Blast Loading. U. S. Army
Engineer Waterways Experiment Station (WES), Structures Laboratory,
Technical Report SL-84-7, Vicksburg, Mississippi.
Sugano, T., Tsubota, H., Kasai, Y., Koshika, N., Orui, S., von Riesemann, W. A.,
Bickel, D. C., and Parks, M. B. (1993). Full-Scale Aircraft Impact Test for
Evaluation of Impact Force. Nuclear Engineering and Design, 140, 373385.
Symonds, P. S., and Mentel, T. J. (1958). Impulsive Loading of Plastic Beams
with Axial Constraints. Journal of the Mechanics and Physics of Solids,
6(3), 186-202.
Symonds, P. S. (1968). Plastic Shear Deformation in Dynamic Load Problems.
Engineering Plasticity, J. Heyman and F. A. Leckie, eds., C.U.P., 647-664.
The Mathworks (2007). MATLAB v7.4 r2007a, Natick, Massachusetts.
Timoshenko, S. P. (1937). Vibration Problems in Engineering. D. Van Nostrand
Company, New York, New York.
U. S. Army Corps of Engineers. (1990). TM5-1300, Structures to Resist the
Effects of Accidental Explosions. U.S. Army, Washington, D.C. (also Navy
NAVFAC P-397 or Air Force AFR 88-22).
Vecchio, F. J., and Collins, M. P. (1986). Modified Compression-Field Theory for
Reinforced Concrete Elements Subjected to Shear. Journal of the
American Concrete Institute, 83(2), 219-231.
Volterra, E., and Zachmanoglou, E. C. (1965). Dynamics of Vibrations. Charles
E. Merril, Columbus, Ohio.
Wen, H. M., Reddy, T. Y., and Reid, S. R. (1995). Deformation and Failure of
Clamped Beams under Low Speed Impact Loading. International Journal
of Impact Engineering, 16(3), 435-454.

235
Wen, H. M., Yu, T. X., and Reddy, T. Y. (1995). Failure Maps of Clamped
Beams under Impulsive Loading. Mechanics of Structures and Machines,
23(4), 453-472.
Wierzbicki, T., and Teng, X. (2003). How the Airplane Wing Cut through the
Exterior Columns of the World Trade Center. International Journal of
Impact Engineering, 28(6), 601-625.
Wood, S. L. (1990). Shear Strength of Low-Rise Reinforced Concrete Walls.
ACI Structural Journal, 87(1), 99-107.
Woodson, S. C. and Baylot, J. T. (1999). Structural Collapse: Quarter-Scale
Model Experiments. U.S. Army Engineer Research and Development
Center, Technical Support Working Group, Technical Report SL-99-8,
Waterways Experiment Station (WES), Vicksburg, Mississippi.
Xue, L., and Wierzbicki, T. (2003). High-Speed Impact of Liquid-Filled Cylinders.
Impact and Crashworthiness Laboratory, MIT, Report No. 108, Cambridge,
Massachusetts.
Yu, T. X., and Chen, F. L. (2000). Further Study of Plastic Shear Failure of
Impulsively Loaded Clamped Beams. International Journal of Impact
Engineering, 24(6), 613-629.
Zhu, F., and Lu, G. (2007). A Review of Blast and Impact of Metallic and
Sandwich Structures. Electronic Journal of Structural Engineering, 7,
Special Issue: Loading on Structures, 92-101.
Zineddin, M., and Krauthammer, T. (2007). Dynamic Response and Behavior of
Reinforced Concrete Slabs under Impact Loading. International Journal of
Impact Engineering, 34, 1517-1534.

APPENDICES

236
Appendix A. Description of Experimental Work

A.1. Introduction
A total of eleven series of small-scale RC beams were tested under static and
dynamic load (CHAPTER 4). The tests were conducted in the Robert L. and
Terry L. Bowen Laboratory for Large-Scale Civil Engineering Research at
Purdue University. Geometry, material properties, and reinforcement ratio of
beams were identical in a given series (Table 4.1). Information about the
material and geometric properties of the beams tested as well as the fabrication,
instrumentation, and testing procedures is presented in this appendix.
A.2. Materials
A.2.1. Concrete
Eleven concrete batches were made: one batch per test series (Table A.1).
Beams and cylinders (for compression and split cylinder tests) were fabricated in
each batch. All concrete used was a small-aggregate concrete with mix
proportions (based on dry weight of materials) of 3.5:1.0:1.5 (coarse ASTM C 33
sand: fine mason sand: Type II Portland cement) and a water/cement ratio of
0.5. Properties of the concrete batches made are listed in Table A.1.
The compressive strength of the concrete (fc) was estimated from tests of 4x8-in.
cylinders capped with neoprene pads. Load was applied with a 600-kip Forney
testing machine at an approximate rate of 0.5-kip/s (40 psi/s). This machine had
a resolution of 100 lbf. The variations of the compressive strength with time for all
the concrete batches made are shown in Figure A.1 to Figure A.4. Individual test
results for fc at 28 days and at ages of tests of beams are listed in Table A.2 for
all the concrete batches. Mean values of fc at ages of tests of beams are listed in
Table A.1. For batches 4 and 6, values of fc at ages of static tests of beams
made from those batches are reported. For batches 1, 3, 5, 7-11, values of fc at

237
ages of impact tests of beams made from those batches are reported. The
compressive strength of the concrete in batch 2 was not measured because
cylinders were defective. The average of fc (at ages of static tests of beams) for
concrete batches 1, 3-6 was used as compressive strength (at age of static tests
of beams) for concrete batch 2. This was done because the compressive
strength of the concrete in batches 1-6 is expected to be similar considering that
these batches: (1) had the same mix proportions, (2) were made within two
weeks, (3) used the same batches of fine and coarse sands, and (4) had the
same casting and curing procedures.
The modulus of elasticity of concrete (Ec) was estimated from the measured
stress-strain response of 6x12-in. cylinders loaded up to about 40% of their
compressive strength. Each cylinder was loaded and unloaded three times after
an initial load cycle during which no measurements were made. This first cycle
was applied to allow the instruments (dial gages) to seat. Deformations of
cylinders were measured every 10 kips (354 psi) of applied load using an ELE
International CT-170 mechanical compressometer with a resolution of 1/20,000in and a gage length of 6-in. Load was applied with a 600-kip Forney testing
machine at an approximate rate of 1 kip/s (35 psi/s) and with a resolution of 100
lbf. Stresses were calculated based on nominal cross-sectional areas. Measured
stress-strain data for concrete batches 7-11 are shown in Figure A.5-Figure A.9.
The modulus of elasticity was obtained as the slope of the line that best fit the
stress-strain data (using the least squares method). Individual test results for Ec
(after impact tests, Table 4.2) for concrete batches 7-11 are listed in Table A.3.
Mean values of Ec (after impact tests) are listed in Table A.1. The variation of Ec
with fc is shown in Figure A.10. The data obtained are represented reasonably
well with the commonly used expression:

Ec = 57,000 f 'c

Eq. A.1

238
where both Ec and fc are in psi units. The maximum deviation of values of Ec on
the curve described by Eq. A.1 from mean measured values of Ec (concrete
batches 7-11) was less than 20% (Figure A.10).
The tensile strength of the concrete was estimated from cylinder splitting tests
using 6x12-in. cylinders that were tested in between thin strips of wood. Load
was applied with a 600-kip Forney testing machine at an approximate rate of 20
kip/min and with a resolution of 100 lbf. Individual test results for the cylinder
splitting strength, fsp (after impact tests, Table 4.2) for concrete batches 7-11 are
listed in Table A.3. Mean values of fsp are listed in Table A.1. The variation of fsp
with fc is shown in Figure A.11. Within the range of these data, the cylinder
splitting strength (fsp) can be approximated as 6.7 f 'c [psi].
A.2.2. Reinforcing Steel
Longitudinal steel used in the test beams was plain No. 12-gage black annealed
wire (0.106-in. nominal diameter, American Steel and Wire, Co.). All wires were
purchased in straight lengths (20-in. long pieces) from Lafayette Tool and Die,
Inc. and were from the same mill heat. Wires used in beam series 1-6 were heat
treated at 800F for 4 hours to have a yield plateau. Wires used in series 7-11
were heat treated at 1100F for 4 hours. A higher temperature was used to
anneal the steel in series 7-11 in order to reduce the yield stress of the wires.
Random samples taken from the wires used in the fabrication of the test
specimens were tested in tension. Before testing, the diameter was measured at
five different points along the length of each wire using a caliper with a precision
of 0.001 in. Table A.4 summarizes measured wire diameters for the reinforcing
steel used in series 1-6. The mean wire diameter was 0.10 in., the standard
deviation was 0.001 in., and the maximum measured deviation from nominal
diameter was 0.008 in. The sample wires were tested in a 120-kip Baldwin
universal testing machine set at a scale (1.2-kip maximum load) with a resolution

239
of 10 lbf. Strain rates were approximately 0.00014/s up to yield and 0.0004/s past
yield. Stress-strain curves for each wire were obtained from stresses calculated
using nominal cross-sectional areas and strains measured with an MTS 634.25E54 extensometer with a 2-in. gage length and maximum nonlinearity of 0.0025 in.
Measured steel properties are summarized in Table A.5. Figure A.12 and Figure
A.13 show representative stress-strain curves obtained for wires used in beam
series 1-6 and 7-11, respectively. The drops in these curves are associated with
rupture of the wires. Tests results showed consistency. Wires used in beam
series 1-6 had upper and lower yield stresses of 124 ksi and 119 ksi,
respectively. For the wires used in beam series 7-11, mean upper and lower yield
stresses were 52 ksi and 49 ksi, respectively, and the mean ultimate stress was
58 ksi. Experimentally inferred values of the modulus of elasticity deviated up to
20% from the typical value of 29,000 ksi.
A.3. Description of Test Specimens
All test specimens were 1-in. deep beams with rectangular cross sections,
reinforced only in the longitudinal direction. Each specimen was mounted on a
setup for testing with its longitudinal axis oriented in the vertical direction.
Rotations and displacements were restrained at the lower end of the beams. At
the upper end, rotations and horizontal displacements were restrained and
vertical displacement unrestrained. Transverse loads (static and dynamic) were
applied at midspan. Each specimen was intended to represent a beam with
continuous ends subjected to a point transverse load at mid-span. Details about
the specimens and their fabrication are given in sections A.3.1 through A.3.3.
Details about the test setup are given in sections A.4.1 through A.4.3.
A.3.1. Dimensions
The nominal dimensions of the test specimens are shown in Figure A.14. The
clear span (L) was 12 in. for beam series 1-10, and 8 in. for series 11. Each end

240
of each beam was clamped to restrain its rotation. The length of the clamped
region, in the direction of the longitudinal axis of the beam, was 3.5 in. The
nominal total length of the specimens was 19 in. in beam series 1-10, and 15 in.
in series 11. The width of the cross sections varied from one test series to the
next. The cross-sectional width was 2 in. for beam series 1-2 and 7-8, 4 in. for
series 3-4 and 9-11, and 6 in. for series 5-6. All cross sections were 1-in. deep.
The effective depth (d) of all specimens was 0.75 in., resulting in shear-span to
effective depth ratios (a/d) of 4 and 2.7 in beam series 1-10 and 11, respectively.
A.3.2. Reinforcement Details
Reinforcement details are presented in Figure A.14. The specimens were not
reinforced in the transverse directions. Longitudinal reinforcement consisted of
continuous top and bottom layers of plain No. 12-gage black annealed wire,
placed at 0.25 in. from the top and bottom (respectively) faces of the beams.
To ensure proper bond between the longitudinal reinforcement and the concrete,
the steel was prepared following a procedure described by Moehle (1980). The
wires were initially cleaned with acetone and toluene. Next, they were sprayed
with 10% (by volume) hydrochloric (muriatic) acid and put to rust in a 100%
relative-humidity (RH) room (fog room) for 72 hours. Then, they were removed
from the fog room and cleaned with water and scouring pads to remove loose
rust particles.
A.3.3. Casting and Curing
Beams were cast using foam-board forms. Cylinders were cast in oiled plastic
forms. Molds for casting of beams and cylinders are shown in Figure A.15.
A 9-ft3 capacity mixer was used for mixing the small-scale concrete. The cement
and sand (coarse and fine) were placed in the mixing bowl and mixed at an
approximate rate of 100 rpm for 2 minutes. Then the water was gradually added

241
to the mix as it was being mixed. The total mix was mixed for about 3 minutes
after all the water was added.
Concrete was placed in all forms (cylinders and beams) in three layers, which
were independently compacted using metal (cylinders) and wooden (beams)
rods (Figure A.16). Exposed concrete surfaces were leveled and trowelled
smooth (Figure A.17). Cylinders were capped and beam molds were covered
with plastic sheets after trowelling.
The test specimens were removed from the forms the day after casting and
wrapped with wet burlap. Sheets of polyethylene were used to prevent the
burlap from drying out (Figure A.18). After seven days of moist curing, burlap
was removed from the specimens and they were kept in the laboratory until they
were tested.
A.4. Test Procedure
A.4.1. Supports
Details on the supports used for all test specimens are presented in Figure A.19.
The supports were designed to restrain both ends of a specimen (over a distance
of 3.5-in at each end) against rotation and displacement in the transverse
directions. Each support consisted of two steel plates, one on each side of the
specimen. One plate had a 1-in. x 6-in. cross section. The other had a 2-in. x 6in. cross section and was fastened to two reaction frames (Figure A.19). These
plates were connected to each other by six ASTM A325 -in. diameter bolts.
Each bolt was tightened to attain a pretension force of approximately 4 kips,
measured as described in section A.4.5. The total clamping force at each end of
each test specimen was approximately 24 kips.

242
In the lower support, two 5/8-in. spacer plates, one on each side of the specimen,
were used to transfer reactions from the specimen to the support plates. In the
upper support, one -in. spacer plate was placed on each side of the specimen.
Between each spacer and the closer support plate, two steel rollers with 3/8-in.
diameters were placed to allow axial deformations. This support condition helped
reduce the increase in beam flexural strength caused by the axial load that is
induced in the specimens as they try to stretch at the initial stages of transverse
loading. Because the surfaces of the beams were not perfectly smooth, layers of
hydrostone approximately 1/8-in. thick were placed between the specimen front
and back faces and the spacers to avoid stress concentration at the ends.
A.4.2. Static Test Setup
Static tests were conducted using the setup shown in Figure A.20. Transverse
load was applied at mid-span by means of a Duff-Norton (Model 3501) worm
gear jack. The load was applied through a 1.5-in.-diameter brass load button
attached to the front face of the specimen. The worm gear jack reacted against
an L6x6x1 angle that was bolted to a W24x55 beam using four ASTM A325 5/8in. diameter bolts. This steel beam was supported by two 2C10x25 double
channels (Figure A.21). Each double channel was anchored to the laboratory
strong floor by two 1-in. diameter high-strength steel threaded rods. C-clamps
were used to attach the beam to the double-channel supports.
A.4.3. Impact Test Setup
Fluid impact tests were conducted using a special setup built at Purdues Bowen
Civil Engineering Laboratory. This experimental system (Figure A.21) is a smallscale adaptation of the system used by Karna et al. (2005). This device can
accelerate both containers filled with fluid and solid objects through a 15-ft long
barrel using compressed helium as a propellant. The barrel was a steel pipe with
an inner diameter of 2.05 in. and a wall thickness of 3/16-in. The pipe was
coupled to a Norgren A1038C-A1 air-pilot-actuated Poppet valve (with a

243
response time of approximately 20 ms). When actuated, the valve opens and lets
the compressed helium stored in a reservoir flow into the barrel, propelling the
projectile. The reservoir was a Swagelok 304 L-HDF8 1-gal stainless steel
double-ended cylinder. The reservoir-valve-barrel assembly rested on top of the
W24x55 beam supported by two 2C10x25 double channels described in section
A.4.2 and Figure A.21.
The projectile used in the tests was a cylindrical aluminum container filled with
water. The container was 3.75-in. long and had an inner diameter of
approximately 2 in. The wall thickness of the container was 0.002 in. The
projectiles were accelerated to speeds of up to 360 miles/hour (160 m/s).
A.4.4. Measurements
The measurements taken during the tests included: transverse load (static tests),
projectile velocity (impact tests), beam deflections, and displacements of a
support plate. A high-speed video camera was used to record the collision of the
containers filled with liquid and the specimens.
Beam deflections were measured at the quarter points and midspan as shown in
Figure A.22. All deflection measurements were made using DC linear variable
differential transformers, LVDTs (section A.4.5). LVDTs 1-3 (Figure A.22) were
located behind the specimen. For impact tests, these LVDTs were protected
against impact of water and concrete debris using a steel angle (Figure A.23a).
Threaded bars connected the cores of the LVDTs and the specimen. In the
impact tests, the threaded bars ran through holes drilled in the steel angle
(Figure A.23b). These bars were attached to the beam using aluminum brackets
either embedded in the concrete or cemented to the back face of the beam with
epoxy (Figure A.23c).

244
Displacements were measured at two points of the smaller plate of the upper
support (Figure A.19c) in static and impact tests in series 7-11 to estimate
support rotations. The locations of the displacement measurements are indicated
in Figure A.22. These measurements were made using LVDTs (section A.4.5).
LVDTs 4 and 5 (Figure A.22) were located behind the upper support plates as
shown in Figure A.24a. Threaded bars connected the cores of the LVDTs and
the smaller plate of the upper support. The threaded bars were attached to the
support plate using aluminum brackets (Figure A.24b-c) cemented to the plate
with epoxy. The threaded bar of LVDT 4 ran through holes drilled in the larger
plate of the upper support (Figure A.19c) and the threaded bar of LVDT 5 ran
below that plate (Figure A.24b-c).
In the impact tests, LVDT 5 and its threaded bar were protected against impact of
water and concrete debris using a steel angle (Figure A.24d). Because of specific
conditions of the experimental setup and the location of LVDT 4 (Figure A.22),
providing protection to this sensor against impact of water and concrete debris
was not easy. No protection was provided to LVDT 4 and its threaded bar in
series 11 which was tested first. As a result, LVDT 4 went out of range in all
impact tests in this series and support rotations could not be estimated. Typical
displacement histories at two points of the support plate (obtained using LVDTs 4
and 5) in series 11 are shown in Figure A.25. Different alternatives to protect
LVDT 4 against impact of water and concrete debris were tried in series 7-10,
with different results. The best protection of LVDT 4 was obtained in series 10
which was tested last. Typical displacement histories at two points of the support
plate (obtained using LVDTs 4 and 5) in series 10 are shown in Figure A.26. It
was difficult to keep LVDT 4 protected in impact tests in which the specimens
disintegrated because of the large amount of flying concrete debris. In these
tests, it was difficult to obtain credible estimates of support rotations.

245
Projectile velocity (or impact velocity), vp, was measured in two different ways:
from signal records of photodetectors being aimed at by laser beams intersecting
the trajectory of the projectile, and from slow motion analysis of the videos
recorded with the high-speed camera during the impact tests. The two methods
were in good agreement (the maximum deviation was 4%; see section 4.3.2).
The laser beams were placed parallel to one another and perpendicular to the
trajectory of the projectile, pointing toward the photodetectors. Signal records
from the photodetectors were obtained in each test (typical records are shown in
Figure A.27 and Figure A.28). The signal records indicated the time at which the
projectile obstructed each laser beam. Impact velocity was computed as the ratio
of the distance between the laser beams and the time interval between the drops
in the signals from the photodetectors. Impact velocity was also computed as the
ratio of the length of the container and the time interval during which each laser
beam remained obstructed.
The high-speed camera was placed to the north of the reaction structure where
the specimen was mounted (Figure A.29). The camera was looking south. The
direction of observation (north to south) was perpendicular to the direction in
which the projectile traveled. Side views of the specimen were obtained. To
increase the clearness of the images recorded by the camera during the impact
tests, the amount of water going past (around) the specimen had to be minimized
so that the side views of the beam were clean. To accomplish this goal,
cardboard or foam-board flaps were placed on both sides of the specimen,
covering the full beam length (Figure A.30). These flaps were typically glued to
the specimen with silicone. Even though the flaps were destroyed by impact,
they delayed the progression of water into the region where high-speed images
were being taken. The high-speed videos recorded during the impact tests were
used to estimate impact velocity following four steps:

246
1. A series of consecutive photographic frames taken right before impact was
extracted from the video.
2. A fixed point of reference was placed on each frame. The location of this
point coincided with the front face of the projectile in the first frame.
3. The number of frames between the first frame and the frame in which the
point of reference was closest to the back face of the projectile was
recorded (N).
4. Impact velocity was computed as the ratio lcan/(Nt), where lcan is the
length of the projectile (3.75 in) and
t is the time interval between two
consecutive frames (1/[number of fps]).
The impact velocity was a controlled variable in the tests. This velocity was
changed from one test to the next and depends on the pressure of helium used
as propellant (section A.4.3). Because this pressure was the variable directly
controlled in the tests, its correlation with outcome projectile velocity had to be
determined. This was done experimentally by carrying out impact tests without
any targets. Projectiles were fired against the strong wall of the laboratory using
different pressures of helium. Projectile velocities were measured at the planned
location of the specimens. The relationship between pressure of helium and
impact velocity obtained is shown in Figure A.31.
A.4.5. Instrumentation
Transverse load in the static tests was measured with a 5-kip Lebow load cell
(Model 3124) mounted on the Duff-Norton worm gear jack that was used. The
pretension force in one of the -in. bolts used to clamp the ends of the test
specimens was measured using a 50-kip Omega load washer (Model LCWD50k). Calibration constants given by the fabricator for both the Lebow and
Omega load cells were checked against readings with a Baldwin/Instron 120-kip
universal testing machine. The maximum errors observed were 20 lbs and 1000
lbs for the 5-kip and 50-kip load cells, respectively.

247
The transverse displacements at the upper end of the specimen were measured
with Lucas-Schaevitz DC LVDTs with a range of 0.05-in. Beam deflections were
measured with Lucas-Schaevitz DC LVDTs with a range of 1-in. These sensors
were calibrated before the tests using a Boeckeler micrometer with a sensitivity
of 0.00002-in. In these calibrations a straight line was fitted to plots of
displacement vs. voltage readings. The maximum deviation of any displacement
reading with respect to this line was 0.0004 in. for the 0.05-in. LVDTs and 0.002
in. for the 1-in. LVDTs. The mid-span deflection of the beams in the static tests
was also measured with a 2-in. Federal dial gage with a precision of 0.001 in.
Impact velocity was measured using two laser pointers and two New Focus
nanosecond photo-detectors (Model 1621) monitored with a National Instruments
(NI) Data Acquisition (DA) card (Model PCI 6259). Videos of the impact tests
were obtained using a Photron APX RS Fastcam high-speed digital video
camera operating between 12,500 and 20,000 frames per second (fps). The APX
RS Fastcam can operate at speeds of up to 250,000 fps.
Voltage signals from the LVDTs and load cells in the static tests were acquired
using a Vishay-Measurements Group Model 5100 scanner. Voltage signals from
the LVDTs and photo-detectors in the impact tests were acquired with a NI PCI
6259 DA card capable of scanning channels simultaneously at a rate of 2.5x106
samples per second. The sampling rate in the impact tests ranged between
60,000-100,000 samples/s.
Because the lasers and photodetectors were monitored at a time interval shorter
than the time between two consecutive frames in the high-speed video (60,000100,000 samples/s vs. 12,500-20,000 frames/s), the impact velocities measured
with the lasers and photodetectors were deemed more reliable than the velocities
estimated from high-speed video.

248
A.4.6. Testing Sequence
Before the beginning of any static or impact test, the specimen was mounted on
the reaction frame with its ends supported as described in Section A.4.1. Next,
LVDTs were placed in the locations indicated in Figure A.22 and were connected
to the test specimen using aluminum brackets on the back face of the specimen.
For the static tests, the next step was to setup a dial gage at mid-span of the
beam and take a reference reading. All sensors were zeroed and continuous
scanning and recording of signals from LVDTs and load cells was then initiated.
Load was applied in small increments (50-100 lbf) up to cracking. After cracking,
small displacement increments (of approximately 0.005 in.) were applied up to
yielding. After yielding, larger displacement increments were applied (0.0200.030 in.). The dial gage was read after every load or displacement increment.
Pictures were taken at the end of each test. The typical duration of a static test
was thirty minutes.
For the impact tests, the next step was to load a projectile into the 15-ft long
barrel. Then, the laser beams were pointed towards the photo-detectors. Next,
the high-speed video camera was set up. This included selecting the position of
the camera, appropriate lighting, operation rate (number of fps), resolution, and
type of trigger, as well as starting recording image continuously. Next, continuous
scanning and recording of signals from LVDTs and photo-detectors was initiated
and immediately afterwards the projectile was fired. Recording of sensor signals
and image was stopped shortly after the impact. Pictures were taken after each
test. The typical duration of an impact test (from firing of the projectile to stopping
sensor and image recording) was less than two seconds.

249

Table A.1 Properties of Concrete Batches


Concrete
Batch

Date of
Casting

Type and Number of Specimens Cast

Average Measured Properties of Concrete

4x8-in.
Cylinders

6x12-in.
Cylinders

Compressive
Strength,
fc(1)
(psi)

Cylinder Splitting
Strength,
fsp(2)

Modulus of
Elasticity,
Ec(2)

9200

---*

---*

---*

---*

(MM/DD/YYYY)

Beams

01/08/2009

---*

01/07/2009

---*

01/09/2009

---*

7630

---*

---*

01/11/2009

---*

8410

---*

---*

01/12/2009

---*

9120

---*

---*

01/15/2009

---*

9860

---*

---*

06/19/2009

10

7040

560

4650

06/22/2009

10

6330

530

4230

06/24/2009

10

6390

530

3950

10

06/25/2009

10

7690

580

4190

11

06/29/2009

10

7720

560

4480

8500

(3)

(psi)

(ksi)

(1)

Average of at least three tests (cylinders). Listed values of fc correspond to ages of static or impact tests (see Table 4.2 and Table A.2).

(2)

Average of two tests (cylinders). Listed values of fsp and Ec were measured after impact tests (see Table 4.2 and Table A.3).

(3)

Compressive strength (fc) was not measured for concrete batch 2 because cylinders were defective. The average of fc in concrete batches 1, 3-6 (at age of
static tests of beams made from these batches) was used.

* No 6x12-in. cylinders were cast. Therefore, tensile strength and modulus of elasticity could not be determined with the equipment available in the laboratory.

249

250

Table A.2 Compressive Strength of Concrete


Date of
Concrete
Compression Tests
Casting
Batch* (MM/DD/YYYY)
at 28 Days
1

01/08/2009

Cylinder
1
2
Mean

Strength
(psi)
6680
6540
6610

Compression Tests at Age of


Static or Impact Tests
Cylinder
1
2
3
4
5
6

Age
(days)
130
130
130
132
132
132

Mean
3

01/09/2009

Cylinder
1
2
Mean

Strength
(psi)
5770
6560
6160

Cylinder
1
2
3
4
5
6

9200
Age
(days)
141
141
141
156
156
156

Strength
(psi)
7870
7370
7620
7810
7680
7460

Age
(days)
104
104
104

Strength
(psi)
8250
8450
8540
8410
Strength
(psi)
8930
9150
9070
9420
8840
9330

Mean
4

01/11/2009

Cylinder
1
2
Mean

Strength
(psi)
6460
6800
6630

Cylinder
1
2
3

7630

Mean
5

01/12/2009

Cylinder
1
2
Mean

Strength
(psi)
7230
6930
7080

Cylinder
1
2
3
4
5
6

Age
(days)
113
113
113
119
119
119

Mean
6

01/15/2009

Cylinder
1
2
Mean

Strength
(psi)
7530
7340
7430

Cylinder
1
2
3

Mean

Strength
(psi)
9330
9110
9380
9370
8780
9250

9120
Age
(days)
100
100
100

Strength
(psi)
9460
10000
10130
9860

* Compressive strength was not measured for concrete batch 2 because cylinders were defective.

251
Table A.2 (continued)
Date of
Concrete
Compression Tests
Casting
Batch (MM/DD/YYYY)
at 28 Days
7

10

11

06/19/2009

06/22/2009

06/24/2009

06/25/2009

06/29/2009

1
2

Strength
(psi)
5740
4910

6010

Mean

5550

Cylinder

1
2

Strength
(psi)
4730
5040

4560

Mean

4780

Cylinder

1
2

Strength
(psi)
4930
4740

5020

Mean

4900

Cylinder

1
2

Strength
(psi)
6010
6100

6280

Mean

6130

Cylinder

1
2

Strength
(psi)
6910
6510

7230

Mean

6880

Cylinder

Compression Tests at
Age of Impact Tests

1
2

Age
(days)
91
91

Strength
(psi)
6050
7240

91

7840

Cylinder

Mean

7040

1
2

Age
(days)
91
91

Strength
(psi)
6250
6640

91

6120

Cylinder

Mean

6330

1
2

Age
(days)
99
99

Strength
(psi)
6180
6610

99

6380

Cylinder

Mean

6390

1
2

Age
(days)
109
109

Strength
(psi)
7580
7620

109

7880

Cylinder

Mean

7690

1
2

Age
(days)
80
80

Strength
(psi)
6880
8150

80

8140

Cylinder

Mean

7720

252
Table A.3 Modulus of Elasticity and Cylinder-Splitting Strength of Concrete
Concrete Date of Casting
(MM/DD/YYYY)
Batch
7

06/19/2009

06/22/2009

06/24/2009

10

06/25/2009

11

06/29/2009

Modulus of Elasticity, Ec
Age
(days)
1
115
2
115
Mean

Ec
(ksi)
5140
4160
4650

Age
(days)
1
113
2
113
Mean

Ec
(ksi)
4290
4170
4230

Age
(days)
1
125
2
125
Mean

Ec
(ksi)
4180
3730
3950

Age
(days)
1
125
2
125
Mean

Ec
(ksi)
4320
4060
4190

Age
(days)
1
121
2
121
Mean

Ec
(ksi)
4400
4550
4480

Cylinder

Cylinder

Cylinder

Cylinder

Cylinder

Cylinder-Splitting Strength, fsp


Age
(days)
140
140
Mean

fsp
(psi)
550
570
560

Age
(days)
137
137
Mean

fsp
(psi)
490
570
530

Age
(days)
135
135
Mean

fsp
(psi)
560
500
530

Age
(days)
134
134
Mean

fsp
(psi)
520
640
580

Age
(days)
130
130
Mean

fsp
(psi)
510
610
560

Cylinder
1
2

Cylinder
1
2

Cylinder
1
2

Cylinder
1
2

Cylinder
1
2

Table A.4 Steel Wire Diameter Measurements Series 1-6


Measured Diameter (in)

Sample
Wire

z= 0

z = 0.25l *

z = 0.5l

z = 0.75l

z=l

1
2
3
4
5
6
7
8

0.099
0.100
0.098
0.098
0.098
0.098
0.099
0.099

0.101
0.100
0.098
0.098
0.099
0.099
0.100
0.101

0.100
0.099
0.100
0.100
0.098
0.099
0.101
0.099

0.102
0.100
0.100
0.101
0.098
0.099
0.099
0.100

0.101
0.098
0.098
0.099
0.098
0.098
0.099
0.099

Mean (rounded) Wire Diameter (in)


Standard Deviation (in)

Average
Diameter
(in)

*z=distance from one end of the wire to the point of diameter measurement; l =wire length (20 in).

0.101
0.099
0.099
0.099
0.099
0.099
0.100
0.100
0.100
0.001

253
Table A.5 Measured Reinforcing Steel Properties
Beam Series
(Annealing
Procedure)
1-6
(800F 4hrs.)

Sample
Wire

Modulus of
Elasticity
(ksi)

Upper
Yield
Stress
(ksi)

Lower
Yield
Stress
(ksi)

Strain at
Onset of
Strain
Hardening*

Ultimate
Stress
(ksi)

1
2
3
4
5
6
7
8

31,580
33,070
31,670
33,550
29,600
33,010
33,860
33,460

125.1
123.7
126.8
123.4
125.3
122.6
126.1
122.9

115.8
120.5
120.1
120.0
120.0
118.7
121.7
118.0

-------------------------

115.8
120.5
120.1
120.0
120.0
118.7
121.7
118.0

32,500

124.5

119.4

----

119.4

1430

1.6

1.8

----

1.8

26,430
34,800
25,910
33,090
31,020
33,600
32,920
28,800
26,420
27,120
28,110
34,340

50.3
50.2
53.5
----**
53.2
53.9
54.3
52.1
52.8
51.2
51.5
51.3

48.5
48.9
51.0
48.3
50.6
49.5
50.4
50.0
50.5
48.9
48.3
49.0

0.065
0.067
0.068
0.058
0.066
0.050
0.075
0.058
0.067
0.067
0.066
0.050

56.9
57.9
58.0
58.0
57.1
59.2
57.2
57.1
59.0
56.9
57.2
57.8

30,200

52.2

49.5

0.063

57.7

3420

1.4

1.0

0.008

0.8

Mean
(rounded)
Standard
Deviation
7-11
(1100F 4hrs.

1
2
3
4
5
6
7
8
9
10
11
12
Mean
(rounded)
Standard
Deviation

* Strain hardening was not observed in the wires used in beam series 1-6.
** No distinct upper yield was observed.

254

Compressive Stress at Failure (psi)

10000

8000

6000

4000

Batch 1
Batch 3

2000

0
0

20

40

60

80

100

120

140

160

Age (days)

Figure A.1 Compressive Strength vs. Time - Concrete Batches 1 and 3

Compressive Stress at Failure (psi)

10000

8000

6000

4000
Batch 4
Batch 5

2000

Batch 6

0
0

20

40

60

80

100

120

140

Age (days)

Figure A.2 Compressive Strength vs. Time - Concrete Batches 4-6

160

255

Compressive Stress at Failure (psi)

10000

8000

6000

4000

Batch 7
Batch 8

2000

0
0

20

40

60

80

100

120

140

160

Age (days)

Figure A.3 Compressive Strength vs. Time - Concrete Batches 7-8

Compressive Stress at Failure (psi)

10000

8000

6000

4000
Batch 9
Batch 10

2000

Batch 11

0
0

20

40

60

80

100

120

140

Age (days)

Figure A.4 Compressive Strength vs. Time - Concrete Batches 9-11

160

256
3200

Stress (psi)

2400

1600

800

Cylinder 1 (1)

Cylinder 1 (2)

Cylinder 1 (3)

Cylinder 2 (1)

Cylinder 2 (2)

Cylinder 2 (3)

0
0.0000

0.0002

0.0004

0.0006

0.0008

Strain

Figure A.5 Initial Stress-Strain Curves for Concrete Cylinders Batch 7

3200

Stress (psi)

2400

1600

800

Cylinder 1 (1)

Cylinder 1 (2)

Cylinder 1 (3)

Cylinder 2 (1)

Cylinder 2 (2)

Cylinder 2 (3)

0
0.0000

0.0002

0.0004

0.0006

0.0008

Strain

Figure A.6 Initial Stress-Strain Curves for Concrete Cylinders Batch 8

257

3200

Stress (psi)

2400

1600

800

Cylinder 1 (1)

Cylinder 1 (2)

Cylinder 1 (3)

Cylinder 2 (1)

Cylinder 2 (2)

Cylinder 2 (3)

0
0.0000

0.0002

0.0004

0.0006

0.0008

Strain

Figure A.7 Initial Stress-Strain Curves for Concrete Cylinders Batch 9

3200

Stress (psi)

2400

1600

800

Cylinder 1 (1)

Cylinder 1 (2)

Cylinder 1 (3)

Cylinder 2 (1)

Cylinder 2 (2)

Cylinder 2 (3)

0
0.0000

0.0002

0.0004

0.0006

0.0008

Strain

Figure A.8 Initial Stress-Strain Curves for Concrete Cylinders Batch 10

258

3200

Stress (psi)

2400

1600

800

Cylinder 1 (1)

Cylinder 1 (2)

Cylinder 1 (3)

Cylinder 2 (1)

Cylinder 2 (2)

Cylinder 2 (3)

0
0.0000

0.0002

0.0004

0.0006

0.0008

Strain

Figure A.9 Initial Stress-Strain Curves for Concrete Cylinders Batch 11

259

E c 57,000 f ' c

Figure A.10 Concrete Modulus of Elasticity vs. Compressive Strength

Cylinder Splitting Strength (psi)

600

500

f sp 6.7 f ' c

400

300

200

100

0
0

2000

4000

6000

8000

Compressive Strength at Age of Impact Tests (psi)

Figure A.11 Concrete Tensile (Cylinder Splitting) vs. Compressive Strength

260

Figure A.12 Representative Stress-Strain Curves for Steel Wires Series 1-6

Figure A.13 Representative Stress-Strain Curves for Steel Wires - Series 7-11

261

0.25
0.25

Beam Series 1-10: L = 12 in


Beam Series 11: L = 8 in
3.5

1.0

1.0
1.0

0.25 (same for all sections)

1.0

0.50

1.0

0.25 (same for all sections)


2.0

2.0

Section A-A - Beam Series 1,7

Section A-A - Beam Series 2,8


(typical)

2.0

1.0
1.0

1.0

0.50

1.0

1.0

0.50

1.0

4.0

4.0

Section A-A - Beam Series 3,9,11

Section A-A - Beam Series 4,10

2.0

2.0

(typical)

0.50

1.0

1.0

1.0

0.50

1.0

1.0

3.5

0.50

6.0

6.0

Section A-A - Beam Series 5

Section A-A - Beam Series 6

Notes: All reinforcement is No. 12-gage black annealed wire ( = 0.106-in., nominal).
All dimensions in inches.

261

Figure A.14 Test Specimen: Nominal Dimensions and Reinforcement Details

262

Figure A.15 Molds Prepared for Casting

Figure A.16 Placing and Compaction of Concrete

263

Figure A.17 Specimen Surfaces after Being Leveled and Smoothed

Figure A.18 Curing of Concrete

264

2C10x25

2C10x25

Support
plate

Support
plate
Reaction
frame

1-in. high-strength
steel threaded rod

Steel roller
3/8-in.
Specimen
A325
3/4-in.

Laboratory
reaction
wall
Laboratory
reaction wall

1/8-in.
hydrostone
layer

1/4-in.
spacer
plate

(c) Specimen Upper Support


1-in. high-strength
steel threaded rod

A325
3/4-in.
Centerline
of specimen
Specimen

1/8-in.
hydrostone
layer

5/8-in.
spacer
plate

Specimen

Support
plate
Support
plate

Support
plate

Laboratory
Laboratory
strong
strongfloor
floor
(a) Support Structure - Elevation

Support
plate

Laboratory
strong floor
(b) Support Structure Front View

(d) Specimen Lower Support

264

Figure A.19 Specimen Supports and Reaction Structure (all dimensions in inches)

265

Figure A.20 Static Test Setup

265

266

2C10x25

Poppet valve

Laboratory
reaction wall
Reservoir

Steel pipe
int.=2.05-in.

Coupler

W24x55
Bottle filled
with helium

Specimen

1-in. high-strength
steel threaded rod

2C10x25

2C10x25

Laboratory strong floor

Figure A.21 Impact Test Setup


266

267

1"

LVDT 4 ( 0.05 in)

5"

3/8"

L/4 - 3/8"

LVDT 5 ( 0.05 in)


LVDT 1 ( 1 in)

L/4

LVDT 2 ( 1 in)
L/4

LVDT 3 ( 1 in)
L/4

Beam Series 1-10: L = 12 in


Beam Series 11: L = 8 in

Figure A.22 Location of Deflection Measurements

268

Figure A.23 Protection of LVDTs against Impact and Attachment to Specimen

Figure A.24 Protection and Attachment of LVDTs at Upper Support Plate

269
0.08
0.06

LVDT4
LVDT5

Displacement (in)

0.04
0.02
0.00
0.00

0.01

0.02

0.03

Time (s)

0.04

-0.02
-0.04
-0.06
-0.08

Figure A.25 Typical Displacement Histories of Upper Support Plate in Series 11

0.08
0.06

LVDT 4
LVDT 5

Displacement (in)

0.04
0.02
0.00
0.00

0.01

0.02

0.03

Time (s)

0.04

-0.02
-0.04
-0.06
-0.08

Figure A.26 Typical Displacement Histories of Upper Support Plate in Series 10

270
0.80

Photodetector Signal (V)

0.60

0.40

0.20

0.00
0.005

0.006

0.007

0.008

0.009

0.010

0.011

Laser/Photodetector 1
Laser/Photodetector 2

-0.20

Time (s)

Figure A.27 Photodetector Signal Records for Measurement of Impact Velocity


Series 1_Test 1

0.80

Photodetector Signal (V)

0.60

0.40

0.20

0.00
0.007

0.008

0.009

0.010

0.011

0.012

0.013

Laser/Photodetector 1
Laser/Photodetector 2

-0.20
Time (s)

Figure A.28 Photodetector Signal Records for Measurement of Impact Velocity


Series 8_Test 2

271

Figure A.29 Position of High-Speed Camera in Impact Tests

Figure A.30 Foam-Board Flaps on Both Sides of Specimen in Impact Tests

272

Projectile Velocity (m/s)

200

150

100

50

0
0

50

100

150

200

250

Pressure of Helium in the Reservoir (psi)

Figure A.31 Empirical Relationship between Pressure of Helium Used as


Propellant and Outcome Velocity of Projectile

300

273
Appendix B. Analysis of Response of Equivalent Elasto-Plastic SDOF System to
Idealized Impact Load using the Newmark-Beta Method

The Mathcad algorithm used to obtain the response to idealized impact load of
the equivalent SDOF systems described in section 5.3.6 is presented here. A
typical SDOF dynamic analysis for one of the specimens that had flexural
deformations in the impact tests is shown (Specimen S8_T1, Table 4.6).

Series-8 Beam
Linear Range

Initial Stiffness

Yield load

Ry := 550lbf
y := 0.014in

Yield deflection

Ry
ke :=
y

4 lbf
ke = 3.9 10
in

ke
k2 :=
1000

Stiffness in the non-linear range:

Dynamic Yield Load


(assuming DIF = 1.5)

Fy := 1.5 Ry

Fy = 825 lbf

Mass

M := 0.0020lbf

Damping Coefficient/Critical Damping Coeff.

DF := 0.02

Circular Frequency

:=

Frequency
Period

lbf
k2 = 39
in

f :=
T :=

in

ke
M

2
1
f

Time Step

t := 0.00001s

Damping Coefficient

c := DF 2 ke M

3 1

= 4.4 10 s

f = 705 s

1
3

T = 1.4 10
T
t

= 142

c = 62.1

kg
s

274
Newmark -Method:
.5

1 M
c
F1

t
2

F4

1
M
1
F2 c 2
t
2

2
1
M
c

F3
2
2 t
t

1
t

1
F5
t

1
F6
2

1
F7
2 t

1
1
F8 2
t
2
2

TEST 1
Impact Load:
Fimpact 994lbf

Load Intensity
td

td 0.0019s
F( n )

190

Load Duration

for i 1 n
Fapplied
i

Fimpact if i t td
0 lbf otherwise

F Fapplied
i
i
F
n 1000

P F( n )

(Load Vector)

Beam Response Calculations:


D( n )

x 0
0

v 0
0

Force 0
0

x 0
a 0
0

for i 1 n
xo x
kact

if x Force
ke if

i 1

Force

k2 otherwise
ke otherwise

i 1

Fy

994

994

994

994

994

994

994

994

994

10

...

lbf

275

(Pi Pi1) + F1vi1 F2ai1

F3 + kact

if x xo < 0
kact ke
x

(Pi Pi1) + F1vi1 F2ai1


F3 + kact

Force Force
i

if

i 1

+ x kact

( Forcei1 < Fy )

( Forcei

> Fy

Tol 0.001
x 2 Tol x

i 1

while

> Tol

i 1

Fy Force
i 1

ke

Fy + k2 x

kact

(Pi Pi1) + F1vi1 F2ai1

x 2

F3 + kact

x x x 2
x x 2
Force Force
i

i 1

+ x kact

F6 a
a F4 x F5 v
i 1
i 1
+ F8 a
v F7 x F6 v
i 1
i 1
x x
i

i 1

v v
i

a a
i

Output
Output
Output

i 1

i 1
i, 0

i, 1

i, 2

+ x
+ v
+ a
t

in
a

Output

i, 3

in
s

Output

i
s

Force
lbf

Force
i 1

276
n 1000

Table D( n )

Displacement Table
k

k 1 n
Resistance Table

k 1

Time Table

k 3

k 0

0.5

0.4

Displacementk0.3
0.29
0.2

0.1

0
210

410

610

810

0.01

Timek

Figure B.1 Calculated Displacement History of Equivalent SDOF System


(Specimen S8_T1)

max( Displacement ) 0.31

max( Resistance ) 836.2 10

min( Displacement ) 0

min( Resistance ) 785.3 10

110

500

Resistancek

500

110

0.2

0.4

0.6

Displacementk

Figure B.2 Resistance-Displacement Relationship for Equivalent SDOF System


(Specimen S8_T1)

277
Appendix C. Method to Estimate Initial Shear Demand in Beams Subjected to
Blast Load using Dynamic Response of an Equivalent Linear SDOF System

A method to estimate maximum shear during the initial phase of response of a


beam to blast load was presented in section 6.4. In this method, the initial
dynamic response of the beam was obtained using an equivalent linear SDOF
system. The SDOF system had a time-dependent stiffness to account for the
change of the deflected shape of the beam in the initial phase (section 6.4.4).
The Newmark-Beta method was used to obtain the dynamic response of the
linear SDOF system. This response was calculated using the Mathcad algorithm
presented below. Using the calculated response of the SDOF system and the
concepts introduced in section 6.4.3 about the variation of the factor with time,
the maximum initial shear in the beam was calculated (section 6.4.2, Eq. 6.15).
This calculation of initial shear in the beam was included in the Mathcad
algorithm presented below. A typical analysis for one of the beam models
subjected to blast load listed in Table 6.2 (Beam ID = 6) is shown.

278

279

280

Figure C.1 Pressure vs. Time Curve for Blast Load (Beam ID = 6 in Table 6.2)

Figure C.2 Total Load applied to the Beam vs. Time

281

Figure C.3 Displacement History of Equivalent Linear SDOF System

Figure C.4 Variation of Factor with Time

282

Figure C.5 Maximum Shear (Reaction Force) in the Initial Phase

283
Appendix D. Numerical Method to Estimate Initial Response and Shear Demand
in Beams Subjected to Blast Load using Timoshenko Beam Theory

The use of the Timoshenko beam theory to obtain initial response and shear
demand in a beam subjected to blast load was presented in section 6.5. A
numerical solution of the differential equations of motion of a beam according to
the Timoshenko theory was implemented using Matlab. The Matlab code used is
presented below. Input data for a typical analysis of one of the beam models
subjected to blast load listed in Table 6.2 (Beam ID = 13) is shown.
clear all
clc
%
% INPUT (units = lbf, in, millisec)
%
% Material Properties
% Modulus of elasticity
E=4e6;
% Shear modulus and shear coefficient (section-shape factor)
G=1.5e6;
ksection=5/6;
% Mass density
ro=2.5e2;
%
% Beam Properties
% Section dimensions (width and depth)
b=12;
hbeam=18;
% Clear span
L=216;
% Section area and moment of inertia
A=b*hbeam;
I=b*hbeam^3/12;
% Angular frequency and period of vibration
omega=2.25*pi^2*(E*I/(ro*A*L^4))^0.5;
Period=2*pi/omega;
%
% Load Parameters (uniform, impulsive, triangular pulse)
% Initial pressure
p0=0;
% Peak pressure
pmax=6000;
% Dwell time
td=2;
% Rise time
tr=td/100;
%

284
% Analysis Parameters
% Number and width of intervals to discretize the grid in x-direction
n=999;
h=L/(2*(n+1));
% Width of interval to discretize the grid in t-direction
k=tr/10;
% Mesh ratio
m=k/h;
% Analysis time
ta=1*td;
% Number of intervals to discretize the grid in t-direction
p=ta/k-1;
%
% CALCULATIONS
%
% Time variation of load
q0=p0*b;
for i=1:n+1
x(i)=i*h;
end
X=[0 x];
for j=1:p+1
if j*k<tr
pressure(j)=(pmax-p0)*j*k/tr;
else
if j*k>td
pressure(j)=0;
else
pressure(j)=(pmax-p0)*(1-(j*k-tr)/(td-tr));
%pressure(j)=p0*(1-(j*k)/td);
end
end
q(j)=pressure(j)*b;
t(j)=j*k;
end
T=[0 t];
%
% Deflection and bending angle at the supports at zero time
Y00=0;
Psi00=0;
%
% Initial Conditions
for i=1:n+1
Y0(i)=0;
Psi0(i)=0;
end
%
% Boundary Conditions
for j=1:p+1
Ysupport(j)=0;
Psisupport(j)=0;
Psimidspan(j)=0;
end
%
%

285
% Initialize variables Y and Psi (n+1 x p+1 matrix of zeros)
for j=1:p+1
for i=1:n
Y(i,j)=0;
Psi(i,j)=0;
end
Y(n+1,j)=0;
Psi(n+1,j)=Psimidspan(j);
end
%
% Recurrence formulas for deflection (Y) and bending angle (Psi)
%
% First time step (j+1=1, t=k=deltat)
Y(1,1)=Y(1,1)+((G*ksection*m^2/ro)*(Y0(2)+Y00)+2*(1G*ksection*m^2/ro)*Y0(1)+k^2*q0/(ro*A)(k*G*ksection*m/(2*ro))*(Psi0(2)-Psi00))/2;
Y(n+1,1)=Y(n+1,1)+((G*ksection*m^2/ro)*2*Y0(n)+2*(1G*ksection*m^2/ro)*Y0(n+1)+k^2*q0/(ro*A)+(k*G*ksection*m/(2*ro))*2*Psi0
(n))/2;
Psi(1,1)=Psi(1,1)+((E*m^2/ro)*(Psi0(2)+Psi00)+(2k^2*G*A*ksection/(ro*I)2*E*m^2/ro)*Psi0(1)+(k*G*A*ksection*m/(2*ro*I))*(Y0(2)-Y00))/2;
for i=2:n
Y(i,1)=Y(i,1)+((G*ksection*m^2/ro)*(Y0(i+1)+Y0(i-1))+2*(1G*ksection*m^2/ro)*Y0(i)+k^2*q0/(ro*A)(k*G*ksection*m/(2*ro))*(Psi0(i+1)-Psi0(i-1)))/2;
Psi(i,1)=Psi(i,1)+((E*m^2/ro)*(Psi0(i+1)+Psi0(i-1))+(2k^2*G*A*ksection/(ro*I)2*E*m^2/ro)*Psi0(i)+(k*G*A*ksection*m/(2*ro*I))*(Y0(i+1)-Y0(i-1)))/2;
end
%
% Second time step (j+1=2, t=2k=2*deltat)
Y(1,2)=Y(1,2)+(G*ksection*m^2/ro)*(Y(2,1)+Ysupport(1))+2*(1G*ksection*m^2/ro)*Y(1,1)-Y0(1)+k^2*q(1)/(ro*A)(k*G*ksection*m/(2*ro))*(Psi(2,1)-Psisupport(1));
Y(n+1,2)=Y(n+1,2)+(G*ksection*m^2/ro)*2*Y(n,1)+2*(1G*ksection*m^2/ro)*Y(n+1,1)Y0(n+1)+k^2*q(1)/(ro*A)+(k*G*ksection*m/(2*ro))*2*Psi(n,1);
Psi(1,2)=Psi(1,2)+(E*m^2/ro)*(Psi(2,1)+Psisupport(1))+(2k^2*G*A*ksection/(ro*I)-2*E*m^2/ro)*Psi(1,1)Psi0(1)+(k*G*A*ksection*m/(2*ro*I))*(Y(2,1)-Ysupport(1));
for i=2:n
Y(i,2)=Y(i,2)+(G*ksection*m^2/ro)*(Y(i+1,1)+Y(i-1,1))+2*(1G*ksection*m^2/ro)*Y(i,1)-Y0(i)+k^2*q(1)/(ro*A)(k*G*ksection*m/(2*ro))*(Psi(i+1,1)-Psi(i-1,1));
Psi(i,2)=Psi(i,2)+(E*m^2/ro)*(Psi(i+1,1)+Psi(i-1,1))+(2k^2*G*A*ksection/(ro*I)-2*E*m^2/ro)*Psi(i,1)Psi0(i)+(k*G*A*ksection*m/(2*ro*I))*(Y(i+1,1)-Y(i-1,1));
end
%
% Remainder of analysis
for j=2:p
Y(1,j+1)=Y(1,j+1)+(G*ksection*m^2/ro)*(Y(2,j)+Ysupport(j))+2*(1G*ksection*m^2/ro)*Y(1,j)-Y(1,j-1)+k^2*q(j)/(ro*A)(k*G*ksection*m/(2*ro))*(Psi(2,j)-Psisupport(j));

286
Y(n+1,j+1)=Y(n+1,j+1)+(G*ksection*m^2/ro)*2*Y(n,j)+2*(1G*ksection*m^2/ro)*Y(n+1,j)-Y(n+1,j1)+k^2*q(j)/(ro*A)+(k*G*ksection*m/(2*ro))*2*Psi(n,j);
Psi(1,j+1)=Psi(1,j+1)+(E*m^2/ro)*(Psi(2,j)+Psisupport(j))+(2k^2*G*A*ksection/(ro*I)-2*E*m^2/ro)*Psi(1,j)-Psi(1,j1)+(k*G*A*ksection*m/(2*ro*I))*(Y(2,j)-Ysupport(j));
for i=2:n
Y(i,j+1)=Y(i,j+1)+(G*ksection*m^2/ro)*(Y(i+1,j)+Y(i-1,j))+2*(1G*ksection*m^2/ro)*Y(i,j)-Y(i,j-1)+k^2*q(j)/(ro*A)(k*G*ksection*m/(2*ro))*(Psi(i+1,j)-Psi(i-1,j));
Psi(i,j+1)=Psi(i,j+1)+(E*m^2/ro)*(Psi(i+1,j)+Psi(i-1,j))+(2k^2*G*A*ksection/(ro*I)-2*E*m^2/ro)*Psi(i,j)-Psi(i,j1)+(k*G*A*ksection*m/(2*ro*I))*(Y(i+1,j)-Y(i-1,j));
end
end
%
% Assembly of matrixes with results for Y and Psi at different x and t
Yfinal=[Y00 Ysupport;transpose(Y0) Y];
Psifinal=[Psi00 Psisupport;transpose(Psi0) Psi];
%
% Shear Force and Moment >> Calculate derivatives w.r.t. x of Y, Psi
for j=1:p+2
for i=2:n+1
dydx(i,j)=(Yfinal(i+1,j)-Yfinal(i-1,j))/(2*h);
dpsidx(i,j)=(Psifinal(i+1,j)-Psifinal(i-1,j))/(2*h);
M(i,j)=-E*I*dpsidx(i,j);
V(i,j)=ksection*A*G*(dydx(i,j)-Psifinal(i,j));
end
end
%
% OUTPUT
%
% Maximum shear at the end section and time of maximum shear
[Vmax tv]=max(V(2,:))
%
% Maximum deflection at time of maximum shear
deltatv=max(Yfinal(:,tv));
%
% Variation of deflection (Y), bending angle (Psi), bending moment (M),
% and shear force (V) along the beam at different times
%
% Y and Psi vs x at t=0, tv(time of maximum shear), ta(end of analysis)
subplot(211),plot(X,Yfinal(:,1),X,Yfinal(:,tv),X,Yfinal(:,p+2))
xlabel('Distance from support (in)')
ylabel('Deflection (in)')
legend('t1','t2','t3')
subplot(212),plot(X,Psifinal(:,1),X,Psifinal(:,tv),X,Psifinal(:,p+2))
xlabel('Distance from support (in)')
ylabel('Bending angle (rad)')
legend('t1','t2','t3')
%
% M and V vs x at t=0, tv(time of maximum shear), ta(end of analysis)
%subplot(211),plot(X(1:n),M(2:n+1,1),X(1:n),M(2:n+1,tv),X(1:n),M(2:n+1,
%p+2))
% xlabel('Distance from support (in)')

287
% ylabel('Moment (lbf-in)')
% legend('t1','t2','t3')
%subplot(212),plot(X(1:n),V(2:n+1,1),X(1:n),V(2:n+1,tv),X(1:n),V(2:n+1,
% p+2))
% xlabel('Distance from support (in)')
% ylabel('Shear (lbf)')
% legend('t1','t2','t3')
%
%
% Mid-span deflection history
% subplot(211),plot(T,Yfinal(n+2,:))
% xlabel('Time (millisec)')
% ylabel('Deflection (in)')
% legend('t1','t2','t3')
%
%
% Deflected Shape at t=tv - Comparison vs LS-DYNA
% XLSDYNA=xlsread('Deflected&V','Defl. Shape','P19:P43');
% YLSDYNA=abs(xlsread('Deflected&V','Defl. Shape','R19:R43'));
% plot(X,Yfinal(:,tv+1),XLSDYNA,YLSDYNA)
% xlabel('Distance from support (in)')
% ylabel('Deflection (in)')
% title('Deflected shape at time of maximum support shear, t=tv')
% legend('Timoshenko beam equations','LS-DYNA',0)

VITA

288

VITA

Oscar A. Ardila-Giraldo was born on August 18, 1980 in Medelln, Colombia,


where he grew up and attended college. He received his B.S. in Civil Engineering
from Universidad Nacional (UNAL) de Colombia at Medelln in 2004. After
graduation, he spent six months as a research assistant in the Structural Stability
Group (in Spanish, Grupo de Estabilidad Estructural, GES) at UNAL, where he
worked under the guidance of Professor Jose D. Aristizbal-Ochoa. Then he
moved to the U.S. to pursue graduate studies at Purdue University (West
Lafayette, Indiana) starting in January, 2005.
The author received his M.S. in Civil Engineering (structural engineering was his
area of concentration) from Purdue University in December, 2006. His advisor
was Professor Mete A. Sozen. He did computational research on response of
beams to impulsive loading, and he was part of the Purdue team that simulated
the aircraft impact on the North Tower of the World Trade Center on September
11, 2001. Upon completion of his M.S., the author enrolled the Ph.D. program in
the School of Civil Engineering (Structures area) at Purdue. His doctoral
dissertation consisted of an experimental and analytical investigation on the initial
response of beams to blast and fluid impact. The investigation focused on the
change in failure mode (from flexure to shear) of beams under rapidly applied
load. His advisor was Professor Santiago Pujol.

Вам также может понравиться