You are on page 1of 242

RELIABILITY ANALYSIS OF BASE SLIDING

OF CONCRETE GRAVITY DAMS SUBJECTED TO EARTHQUAKES

Tomas Horyna
Dipl. Engineer, The Czech Technical University in Prague, 1989
M.A.Sc, The University of British Columbia, 1995
A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF
THE REQUIREMENTS FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY
in
THE FACULTY OF GRADUATE STUDIES
(Department of Civil Engineering)
We accept this thesis as conforming
to thej^uij?erJ~standaid

THE UNIVERSITY OF BRITISH COLUMBIA


AUGUST 1999
Tomas Horyna, 1999

In presenting this thesis in partial fulfilment of the requirements for an advanced


degree at the University of British Columbia, I agree that the Library shall make it
freely available for reference and study. I further agree that permission for extensive
copying of this thesis for scholarly purposes may be granted by the head of my
department

or by

his or

her representatives.

It

is understood

that

copying

or

publication of this thesis for financial gain shall not be allowed without my written
permission.

Department of

CIVIL EWGIMEEPIN^

The University of British Columbia


Vancouver, Canada
Date

DE-6 (2/88)

AUGUST Sjhf WW

ABSTRACT
Concrete gravity dams are typically constructed in blocks separated by vertical contraction
joints. The design of straight concrete gravity dams is traditionally performed by assuming each
block to be independent, except for gravity dams in valleys with relatively small width to height
ratios. Understanding the 2-D behaviour of individual monoliths is thus considered relevant and
2-D models are usually employed in safety evaluations of existing dams.
During a strong seismic event, low to medium height concrete gravity dams tend to crack at the
base as opposed to tall dams, which attract high stresses and cracking at the level of a slope
change on the downstream side of a dam. The state-of-the-practice in the seismic evaluation of
concrete gravity dams requires that the failure mode of the dam monolith sliding at its base be
considered.
This study focused on the post-crack dynamic response of existing concrete gravity dams in
order to investigate their safety against sliding considering non-linear effects in the damfoundation interface. Sliding response of a single monolith of a low to medium height concrete
gravity dam at the failure state was studied and, therefore, the monolith separated or unbonded
from its foundation was considered. The work included experimental, analytical and reliability
studies.
During the experimental study, a model of an unbonded concrete gravity dam monolith was
developed and tested using a shake table. The model, preloaded by a simulated hydrostatic
force, was subjected to a selected variety of base excitations.

Other effects, such as

hydrodynamic and uplift pressures were not considered in the experiments. A strong influence
of amplitude and frequency of the base motions on the sliding response of the model was
observed during the tests.

ii

ABSTRACT

Simple and more detailed numerical models to simulate the experiments were developed during
the analytical study. It was observed that a simple rigid model could simulate acceptably the
tests only in a limited range of excitation frequencies. A finite element (FE) model simulated
the experiments satisfactorily over a wider range of dominant frequencies of the base
accelerations.
The numerical models were used to simulate the seismic response of a 45 m high monolith of
a concrete gravity dam subjected to three different earthquake excitations for varying
reservoir's water level. The agreement between the results using the simple rigid and the F E
models was found acceptable.
The results of the numerical simulations were used in a reliability analysis to calculate
probabilities of failure of the 45 m high monolith. Probability of failure was defined here as an
annual chance of exceeding an allowable amount of the monolith's base sliding during an
earthquake. The peak ground acceleration (PGA), the characteristics of the time history, and the
reservoir's water level were considered as random parameters during this study. Using the F E
model, the annual probabilities of failure ranged from 1. 1E-8 for the mean P G A of 0.2g and 20
cm of allowable sliding to 1.3E-3 for the mean P G A of 0.6g and 1 cm of allowable sliding. The
probabilities of failure using the simple rigid model were found close to those using the F E
model. It was concluded that the computationally less demanding simple rigid model may be
adequately used in reliability calculations of low to medium height concrete gravity dam safety
against base sliding.

T A B L E OF CONTENTS
ABSTRACT

ii

T A B L E OF CONTENTS

iv

LIST OF T A B L E S

. . . . ix

LIST OF FIGURES

LIST OF S Y M B O L S A N D A B B R E V I A T I O N S

xiv

ACKNOWLEDGEMENTS
DEDICATION

xvi
-

CHAPTER 1. INTRODUCTION

xvii

1.1. G E N E R A L

1.2. OBJECTIVES A N D SCOPE OF THIS S T U D Y

1.3. THESIS OUTLINE

CHAPTER 2. L I T E R A T U R E R E V I E W
2.1. C O N C R E T E D A M S DURING PAST E A R T H Q U A K E S

8
9

2.2. E X P E R I M E N T S O N A C T U A L C O N C R E T E G R A V I T Y D A M S

11

2.3. E X P E R I M E N T S O N S C A L E D M O D E L S

11

2.3.1. Niwa and Clough (1980)

12

2.3.2. Donlon and Hall (1991)

12

2.3.3. Zadnik and Paskalov (1992)

12

2.3.4. L i n et al. (1993)

13

2.3.5. Mir and Taylor (1995)

13

2.3.6. Mir and Taylor (1996)

14

2.3.7. Tinawi et al. (1998 b, c)

14

2.4. TRADITIONAL SEISMIC A N A L Y S I S OF C O N C R E T E


GRAVITY DAMS
2.5. REFINED SEISMIC A N A L Y S I S OF C O N C R E T E G R A V I T Y D A M S

15
17

2.6. SEISMIC A N A L Y S I S RECOGNIZING E X P L I C I T L Y


THE DAM-FOUNDATION CONTACT PLANE
2.6.1. Leger and Katsouli (1989)

20
20
iv

TABLE OF CONTENTS

2.6.2. Chopra and Zhang (1991)

21

2.6.3. Danay and Adeghe (1993)

21

2.6.4. Chavez and Fenves (1995 and 1996)

21

2.6.5. Mir and Taylor (1996)

24

2.6.6. Tinawi et al. (1998b, c)

24

CHAPTER 3. P R E L I M I N A R Y E X P E R I M E N T A L STUDY OF A D A M M O D E L . . . . 26
3.1. OBJECTIVES A N D SCOPE OF P R E L I M I N A R Y TESTS

26

3.2. DESCRIPTION OF TESTING FACILITY

27

3.3. E X P E R I M E N T A L M O D E L

29

3.3.1. Shape of the Model for a Dam Monolith

29

3.3.2. Material for Monolith and Similitude Considerations

31

3.3.3. Sliding Surfaces

34

3.3.3.1. Smooth Surface

36

3.3.3.2. Cracked-ice surface

37

3.3.3.3. Rough surface

38

3.3.4. Simulation of Hydrostatic Load

39

3.4. E X P E R I M E N T A L SETUP A N D TESTING PROCEDURES

39

3.4.1. Static Tests

40

3.4.2. Dynamic Tests

42

3.5. INSTRUMENTATION

45

3.6. S U M M A R Y A N D CONCLUSIONS

46

CHAPTER

4.

F U R T H E R E X P E R I M E N T A L STUDY OF THE D A M M O D E L

49

4.1. OBJECTIVES A N D SCOPE OF FURTHER EXPERIMENTS

49

4.2. E X P E R I M E N T A L M O D E L

50

4.2.1. Sliding Surfaces

50

4.2.2. Hydrostatic Load Simulation

52

4.3. INSTRUMENTATION

54

4.4. TEST SETUP A N D PROCEDURES

56
v

TABLE OF CONTENTS

4.4.1. Impact Hammer Tests

56

4.4.2. Static Tests

58

4.4.3. Tests with Harmonic Input

61

4.4.4. Tests with Synthetic Earthquake Input

62

4.5. RESULTS OF EXPERIMENTS

65

4.5.1. Results of Impact Hammer Tests

66

4.5.2. Results of Static Tests

68

4.5.3. Results of Tests with Harmonic Input

71

4.5.4. Results of Tests with Synthetic Earthquake Input

74

4.5.5. Comparison of Results from Harmonic and Earthquake Tests

80

4.6. S U M M A R Y OF E X P E R I M E N T A L WORK.

83

CHAPTER 5. A N A L Y T I C A L STUDY
OF THE D A M M O N O L I T H M O D E L
5.1. D E V E L O P M E N T OF N U M E R I C A L M O D E L S

85
85

5.1.1. SDOF Numerical Model

86

5.1.2. 3DOF Numerical Model

94

5.1.3. Finite Element Model

96

5.1.3.1. Previous F E Modeling at U B C

96

5.1.3.2. A N S Y S 5.3 Multiphysics/Universify

97

5.1.3.3. Description of the Model

97

5.1.3.4. Plane Element

98

5.1.3.5. Contact Element

99

5.2. C A L I B R A T I O N OF THE F E M O D E L

100

5.2.1. Modifications of the Original FE Model

100

5.2.2. Modal Analysis

101

5.3. RESULTS F R O M SIMULATIONS OF S H A K E T A B L E TESTS

107

5.3.1. Parameters of Simulations

107

5.3.2. Comparison of Amounts of Sliding

109

5.3.2.1. Simulations of Harmonic Tests

109

vi

TABLE OF CONTENTS

5.3.2.2. Simulations of Synthetic Earthquake (EQ2) Tests

111

5.3.2.3. Summary of Comparisons

111

5.4. A N A L Y S I S OF SLIDING/ROCKING OF 3DOF M O D E L

115

5.5. S U M M A R Y

118

CHAPTER

6.

A N A L Y T I C A L STUDY OF
A FULL-SCALE D A M MONOLITH

120

6.1. OBJECTIVES A N D SCOPE OF THE A N A L Y T I C A L S T U D Y


ON A F U L L - S C A L E MODEL

120

6.2. DESCRIPTION OF THE STRUCTURE

121

6.3. DESCRIPTION OF B A S E EXCITATIONS

124

6.4. FINITE E L E M E N T STUDY

125

6.4.1. Description of the FE Model for the Dam Monolith

125

6.4.2. Modal Characteristics of the Dam Monolith

130

6.4.3. Nonlinear Time History Analysis Using FE Model

136

6.5. S T U D Y USING SDOF M O D E L

137

6.5.1. Description of SDOF Model for the Dam Monolith

137

6.5.2. Nonlinear Time History Analysis Using SDOF Model

138

6.6. C O M P A R I S O N OF RESULTS F R O M THE FE


A N D SDOF SIMULATIONS
CHAPTER

7.

138

R E L I A B I L I T Y S T U D Y OF
A FULL-SCALE D A M MONOLITH

143

7.1. BASIC CONCEPTS OF RELIABILITY A N A L Y S I S

143

7.2. OBJECTIVES A N D SCOPE OF R E L I A B I L I T Y A N A L Y S I S

147

7.3. R E D U C T I O N OF THE A N A L Y T I C A L RESULTS

148

7.4. R E L I A B I L I T Y S T U D Y WITH TWO R A N D O M V A R I A B L E S

151

7.4.1. Peak Ground Acceleration as a Random Variable

152

7.4.2. Effect of Different Earthquake Records Treated as


an Error Random Variable

156

vii

TABLE OF CONTENTS

7.4.3. Simplified Reliability Analysis

157

7.4.4. Results of Reliability Analysis with Two Random Variables

159

7.5. R E L I A B I L I T Y S T U D Y WITH THREE R A N D O M V A R I A B L E S

163

7.6. RESULTS OF S T U D Y IN TERMS OF S A F E T Y FACTORS

168

7.7. I N F L U E N C E OF FOUNDATION PROPERTIES


O N THE PROBABILITIES OF F A I L U R E
7.8. S U M M A R Y
CHAPTER 8.

S U M M A R Y , CONCLUSIONS, A N D R E C O M M E N D A T I O N S

8.1. S U M M A R Y A N D CONCLUSIONS

172
175
178
178

8.1.1. Experimental Study

178

8.1.2. Analytical Study

179

8.1.3. Reliability Study

181

8.2. R E C O M M E N D A T I O N S FOR FURTHER R E S E A R C H


REFERENCES

183
185

A P P E N D I X A - PLOTS OF M O D E L MOTIONS F R O M
NO-SLIP TESTS A N D SIMULATIONS

190

APPENDIX B - S A M P L E S OF INPUT FILES


FOR A N S Y S A N A L Y S I S

201

A P P E N D I X C - PLOTS F R O M F E A N D SDOF SIMULATIONS


ON A FULL-SCALE D A M MONOLITH

208

A P P E N D I X D - C A L C U L A T I O N S FOR RELIABILITY STUDY


USING RESULTS OF F E SIMULATIONS

215

A P P E N D I X E - C A L C U L A T I O N S FOR R E L I A B I L I T Y S T U D Y USING R E S U L T S OF
SDOF SIMULATIONS

220

viii

LIST OF TABLES
Table 4.1: List of Instrumentation and Location

56

Table 4.2: Natural Frequencies of the Bonded Model

67

Table 4.3: Natural Frequencies of the Unbonded Model

67

Table 4.4: Friction Coefficients of Rough Surface R l Obtained from Static Tests

70

Table 4.5: Grouping Testing Frequencies According to Response of the Model

81

Table 5.1: Natural Frequencies of the Experimental Model Bonded to the Base

102

Table 5.2: Natural Frequencies of the Unbonded Experimental Model

102

Table 6.1: Natural Frequencies of the Dam Monolith Bonded to the Base

131

Table 6.2: Natural Frequencies of the Unbonded Dam Monolith

131

Table 7.1: Example of Results for a site with a =0.4g


475

and for an Allowable Sliding of 5 cm


Table 7.2: Safety Factors Evaluated for Parameters Used in Reliability Study

167
170

Table 7.3: Safety Factors Evaluated for the Target Probability of 1E-5
Based on A E P of 1/475

171

Table 7.4: Example of Results for a site with a =0.4g


475

and for an Allowable Sliding of 5 cm


Table 7.5: Results of FE Analyses with Varied Friction Coefficient

174
175

ix

LIST OF FIGURES
Figure 2.1: Model of the Dam-Water-Foundation System Used in Program E A G D - S L I D E 22
Figure 3.1: Main Parts of Experimental Model

30

Figure 3.2: Schematic of the Model for a Dam Monolith

31

Figure 3.3: Dimensions of the Upper Surface Plate

35

Figure 3.4: Dimensions of the Lower Surface Plate

36

Figure 3.5: Connection of USP to the Monolith

37

Figure 3.6: Steel Frame for the LSP to Shake Table Connection

38

Figure 3.7: Controlling the Level of Hydrostatic Force during a Dynamic Test

40

Figure 3.8: Experimental Setup for Static Tests

41

Figure 3.9: Photo of Experimental Setup for Static Tests

41

Figure 3.10: Experimental Setup for Dynamic Tests

43

Figure 3.11: Photo of Experimental Setup for Dynamic Tests

43

Figure 3.12: Displacement Sensors Located on Downstream Side of the Monolith

45

Figure 3.13: View of a Load Cell to Measure Force between the Model and the Rigid Arm 46
Figure 4.1: Schematic of the Upper Surface Plate

51

Figure 4.2: Photo Showing High Pressure Jet Preparing Rough Surface

52

Figure 4.3: Schematic of Hydrostatic Load Assembly

53

Figure 4.4: Photo of the Model and the Hydrostatic Load Simulation Assembly

53

Figure 4.5: Location of Instrumentation

55

Figure 4.6: Instrumented Model During Hammer Testing

57

Figure 4.7: View of Downstream Side of the Model Attached to Rigid Floor
during Impact Hammer Tests

57

Figure 4.8: Schematic of Strong Arm Used For Static Tests

59

Figure 4.9: Photo of Typical Static Test Setup

59

Figure 4.10: Winch Assembly Used to Pull Model Up-Stream Between Tests

61

Figure 4.11: Example of Harmonic Input Record at 5 Hz

62

Figure 4.12: Typical Harmonic Test Setup

63

Figure 4.13: Example of Synthetic Earthquake Input Record


with Dominant Frequency 5 Hz

65
x

LIST OF FIGURES

Figure 4.14: Example of Signals Recorded during Static Tests with Surface RI

69

Figure 4.15: Example of Signals Recorded during Harmonic Tests with Surface RI

72

Figure 4.16: Table Acceleration vs. Rate of Model Displacement for Harmonic Excitation 73
Figure 4.17: Example of Signals Recorded during Earthquake Tests with Surface R I . . . . 76
Figure 4.18: Table Acceleration vs. Rate of Model Displacement for Excitation EQ1 . . . . 77
Figure 4.19: Table Acceleration vs. Rate of Model Displacement for Excitation EQ2 . . . . 78
Figure 4.20: Table Acceleration vs. Rate of Model Displacement for Excitation EQ3 . . . . 79
Figure 5.1: Schematic of the SDOF Model

87

Figure 5.2: Performance of SDOF Model - Harmonic Excitations at 5 H z and 0.4g

90

Figure 5.3: Performance of SDOF Model - Harmonic Excitations at 5 Hz and 0.6g

91

Figure 5.4: Performance of SDOF Model - Harmonic Excitations at 5 Hz and l . l g

92

Figure 5.5: Performance of SDOF Model - Harmonic Excitations at 5 Hz and 1.4g

93

Figure 5.6: Comparison of Results from 3DOF and SDOF Models


for Harmonic Excitations

95

Figure 5.7: Schematic of the FE Model

98

Figure 5.8: F E Model for Experimental Model of Dam Monolith Bonded to Base

103

Figure 5.9: First Natural Mode of Experimental Model Bonded to Base; f=65.8 Hz

103

Figure 5.10: Second Natural Mode of Experimental Model Bonded to Base; f=l15.4 H z . 104
Figure 5.11: Third Natural Mode of Experimental Model Bonded to Base; f=173.5 H z . . 104
Figure 5.12: F E Model for Unbonded Experimental Model of Dam Monolith

105

Figure 5.13: First Natural Mode of Unbonded Experimental Model; f=29.7 Hz

105

Figure 5.14: Second Natural Mode of Unbonded Experimental Model; f=48.9 Hz

106

Figure 5.15: Third Natural Mode of Unbonded Experimental Model; f= 143.3 Hz

106

Figure 5.16: Comparison of Sliding from Tests and Simulations for Harmonic Excitations 110
Figure 5.17: Comparison of Sliding from Tests and Simulations for EQ2 Excitations . . . 112
Figure 5.18: Kinematics of the Dam Monolith Model

116

Figure 6.1: Schematic of the Full-Scale Monolith Structure

122

Figure 6.2: Characteristics of Selected Record from 1985 Nahanni Earthquake

126

Figure 6.3: Characteristics of Selected Record from 1985 Mexico Earthquake

127
xi

LIST OF FIGURES

Figure 6.4: Characteristics of Selected Record from 1994 Northridge Earthquake

128

Figure 6.5: F E Model for a Full-Scale Dam Monolith Bonded to Base

132

Figure

6.6: First Natural Mode of a Full-Scale Dam Monolith Bonded to Base; f=6.9 Hz. 132

Figure 6.7: Second Natural Mode of a Full-Scale Dam Monolith


Bonded to Base;

133

f=16.9 Hz

Figure 6.8: Third Natural Mode of a Full-Scale Dam Monolith


Bonded to Base;

133

f=18.8 Hz

Figure 6.9: F E Model for an Unbonded Full-Scale Dam Monolith

134

Figure 6.10:

First Natural Mode of an Unbonded Full-Scale Dam Monolith; f=4.3 H z . . .

134

Figure 6.11:

Second Natural Mode of an Unbonded Full-Scale Dam Monolith; f=9.3 Hz.

135

Figure

6.12: Third Natural Mode of an Unbonded Full-Scale Dam Monolith; f=16.2 Hz . 135

Figure 6.13:

Dam Monolith Displacements (Sliding) vs. Reservoir Water Level

140

Figure 6.14:

Dam Monolith Displacements (Sliding) vs. Peak Ground Acceleration . . . .

141

Figure 7.1: Values of P G A Necessary to Cause Sliding of 1 mm

155

Figure 7.2: Annual Probabilities of Failure vs. Water Level


of the Reservoir for A E P of 1/475

161

Figure 7.3: Annual Probabilities of Failure versus Acceleration

165

Figure 7.4: Forces on the interface zone of the gravity dam monolith

169

Figure 7.5: Safety Factors for a Target Annual Probability of 1E-5


Based on A E P of 1/475

172

Figure A.1: Response of Dam Model to Harmonic Excitation at Frequency of 5 H z

192

Figure A.2: Response of Dam Model to Harmonic Excitation at Frequency of 7.5 Hz . . . 193
Figure A.3: Response of Dam Model to Harmonic Excitation at Frequency of 10 H z . . . .

194

Figure A.4: Response of Dam Model to Harmonic Excitation at Frequency of 12.5 Hz . .

195

Figure A.5: Response of Dam Model to Harmonic Excitation at Frequency of 15 H z . . . .

196

Figure A.6: Response of Dam Model to Harmonic Excitation at Frequency of 17.5 Hz . .

197

Figure A.7: Response of Dam Model to Harmonic Excitation at Frequency of 20 H z . . . .

198

Figure A.8: Response of Dam Model to Harmonic Excitation at Frequency of 22.5

Hz . .

199

Figure A.9: Response of Dam Model to Harmonic Excitation at Frequency of 25 H z . . . .

200
xii

LIST OF FIGURES

Figure C.1: Response of Dam Monolith to Nahanni Earthquake Record with PGA 1.04g. 209
Figure C.2; Response of Dam Monolith to Nahanni Earthquake Record with PGA 0.79g. 210
Figure C.3: Response of Dam Monolith to Mexico Earthquake Record with PGA 1.04g . 211
Figure C.4: Response of Dam Monolith to Mexico Earthquake Record with PGA 0.79g .212
Figure C.5: Response of Dam Monolith to Northridge Earthquake Record
with PGA 1.04g

213

Figure C.6: Response of Dam Monolith to Northridge Earthquake Record


with PGA 0.79g

214

Figure D.1: Sliding vs. PGA Relationships for Different Water Levels Using Results
of FE Analyses

216

Figure D.2: Linear Regression of 'a' coefficients for FE Analysis Results

217

Figure D.3: Linear Regression of'b' coefficients for FE Analysis Results

217

Figure D.4: Comparison of Sliding Obtained from FE Simulations


and Values Calculated Using Interpolation Function

218

Figure D.5: Distribution of Errors due to Different Earthquake Records vs. PGA's

218

Figure D.6: Annual Probabilities of Exceeding a Given Amount of Sliding


from Simplified Analysis Using FE Results

219

Figure E.1: Sliding vs. PGA Relationships for Different Water Levels Using Results
of SDOF Analyses

.221

Figure E.2: Linear Regression of'a' coefficients for SDOF Analysis Results

222

Figure E.3: Linear Regression of'b' coefficients for SDOF Analysis Results

222

Figure E.4: Comparison of Sliding Obtained from SDOF Simulations


and Values Calculated Using Interpolation Function

223

Figure E.5: Distribution of Errors due to Different Earthquake Records vs. PGA's

223

Figure E.6: Annual Probabilities of Exceeding a Given Amount of Sliding


from Simplified Analysis Using SDOF Results

224

xiii

LIST OF SYMBOLS AND ABBREVIATIONS


3 DOF =
A =

3 Degrees Of Freedom
constant

a, a\, a = constants
2

AEP =

Annual Exceedence Probability

AEP

P k ground acceleration corresponding to a given A E P

e a

= mean value of design acceleration

B=
b, b\,b =
2

Peak Ground Acceleration

constant
constants

C = capacity of a system, cohesion, constant


CEA =
CSMIP =

Canadian Electricity Association


California Strong Motion Instrumentation Program

D = demand on a system
d/s =
E =
FE =

downstream
constant
Finite Elements

FS = Factor of Safety
g = acceleration due to gravity
G = performance function
H=
LVDT =

m, m

depth of the reservoir


Linearly Varying Displacement Transformer

= mass, added mass of water

HD

MDE -

Maximum Design Earthquake

Mff = total added mass of water


D

N =

normal force

P and P = annual and event probability of failure


a

Pf = probability of failure
P G A = Peak Ground Acceleration (in text)
P=
r

7? =
N

reliability (probability of survival)


Standard Normal variable
xiv

LIST OF SYMBOLS

AND

ABBREVIATIONS

S = sliding of a gravity dam monolith or the experimental model


S =
0

SDOF=

allowable (limiting) amount of sliding


Single Degree Of Freedom

u/s = upstream
USGS =

U.S. Geological Survey

V=

coefficient of variation, dynamic tangential force

V =

coefficient of variation, dynamic tangential force

st

vs. = versus
WM =

Working Model (software)

P = reliability index
Y =

specific mass of water

p =

friction coefficient

Q> = Standard Normal probability distribution function

xv

ACKNOWLEDGEMENTS
This project was conducted with research funding from the British Columbia Hydro and Power
Authority and from the Natural Sciences and Engineering Research Council of Canada. Support
of both institutions is gratefully acknowledged.
There have been many people who have helped me during the various tasks associated with the
completion of this thesis. Their assistance is very much appreciated, and I would not have been
able to accomplish as much without their assistance.
I would like to thank my supervisors, Dr. Ricardo Foschi and Dr. Carlos Ventura, for the
guidance and technical advice they have given me over the years. I am personally indebted to
Dr. Ricardo Foschi for financing my graduate studies at U B C . I want to thank both my
supervisors for giving me the opportunity to work with them on a number of interesting research
and educational projects and for their personal advice and assistance during all those years.
Finally, their comments and suggestions during the preparation of this thesis are sincerely
appreciated.
Professional advice and valuable comments in various stages of the project from several B C
Hydro structural engineers is greatly acknowledged. These included Mr. Benedict Fan, Mr.
Gilbert Shaw, Dr. Desmond Hartford and others.
The experimental section of this thesis would not have been possible without the assistance of
the Laboratory technicians. Special thanks go to Howard Nicol for his valuable hints for the
testing, and to Doug Hudniuk for his machining and welding help with the experimental model.
Several U B C graduate students contributed to the successful completion of this project. Among
all these, Jachym Rudolf and Cameron Black contributed the most. Vincent Latendresse, Hong
L i , Hugo Armelin and Mahmoud Rezai assisted during the building of the experimental model
and shake table testing.
I would like to thank Dr. D. L . Anderson of U B C and Mr. B. H . Fan of B C Hydro, who, along
with Dr. Foschi and Dr. Ventura, reviewed this thesis.
Last, but certainly not the least, I would like to thank my wife Petra and our family. Petra is
given special thanks for the patience, care and encouragement she has had for me during the
years I spent studying and working at U B C .

xvi

To Petra and our parents

INTRODUCTION

1.1 GENERAL
Failures of dams are very infrequent, but in most cases, these would be extremely high
consequence events. Therefore, the assessment of dam safety is treated with the utmost care.
The good performance of concrete gravity dams in past earthquakes has been used by some to
support their claim that these structures are inherently strong in withstanding the effects of
earthquakes.

Others maintain that this apparently good behaviour should be viewed in the

proper perspective: that recorded incidents may not have involved dams experiencing the most
severe earthquakes assessed possible for their respective sites and with full reservoirs, and that
performance data reporting may not have been complete (Fan and Sled, 1992).
For many dams built at the beginning of this century, the design as well as the construction
specifications did not correspond to today's requirements. It is not surprising if many existing
dams initially considered safe are now judged unsafe based on up-to date specifications (Tinawi
etal., 1998b).
Concrete gravity dams are typically constructed in blocks separated by vertical contraction and
horizontal construction joints. The vertical joints, extending from the foundation to the crest and
from the upstream to the downstream faces, are used to limit the cross-canyon dimension of the
blocks, called monoliths.

The horizontal joints are provided to divide the structure into

convenient building units and to regulate temperature during concrete placement.

It is

reasonable to assume that joints have a marked weakening effect on concrete gravity dams. This
weakening depends, mostly, on the low capacity of the joints to transfer loads when compared
to mass concrete.

Therefore, the mechanical properties of joints in gravity dams reduces


1

CHAPTER

INTRODUCTION

significantly their ability to resist severe earthquakes and they play an important role in
assessment of dam safety.
The design of concrete gravity dams is generally performed by assuming complete bonding in
the horizontal joints. However, for a detailed safety evaluation of an existing dam, it is
necessary to characterise properly these joints, evaluate the possibility of relative motions, such
as sliding or separation, which influence the stability of the structure. Therefore, it is necessary
to consider these interfaces with proper characteristics in dam safety analysis. A logical location
of one of the important horizontal or near horizontal joints is the interface between the dam body
and the foundation rock. Possibility of sliding in this interface has to be evaluated during safety
analysis of a concrete gravity dam.
The need to accurately predict the seismic stability of concrete gravity dams has led to a
proliferation of research activities. A large number of analytical methodologies have been put
forward. Material strength parameters, ground motion characteristics and failure mechanisms
are represented with varying degrees of rigor (Fan and Sled, 1992). These procedures are
expensive and time consuming, and they produce results that are generally only snapshots of the
dam performance. Certain procedures, with simplifying assumptions, enable the prediction of
the amount of earthquake induced sliding of gravity dams.

Such analyses do provide

information on damage caused during earthquakes, which is useful the for evaluation of postearthquake safety of the dam in its damaged state. They do not, however, by themselves enable
a credible assessment of dam safety during the earthquake, unless a correlation could be made
between the computed damage and probability of dam failure.

CHAPTER

INTRODUCTION

A n alternative to assessing the seismic safety of dams is to assess the existing seismic risks and
compare them to acceptable risks. Possible failure mechanisms are first screened and probable
ones identified. A range of ground motion characteristics, reservoir conditions, and material
strengths are considered. For each combination of parameters, each of the identified probable
failure mechanism is checked for its stability. A data set on the fragility of the dam is thus
constructed. A relatively large number of such stability checks would have to be made, and it
becomes readily apparent that a simple numerical procedure is preferred.
Current state-of-the-art analytical tools to evaluate sliding of concrete gravity dams are rarely
used for seismic risk evaluations because of time constraints, but they can yield valuable results
to verify simpler methods.

Experimental data is another source of verification of simple

analytical methods. As there is no reported evidence of excessive sliding of any actual concrete
gravity dam subjected to severe shaking, it is desirable to conduct shake table tests of dam
models to verify analytical predictions.
Concrete gravity dams represent about half of British Columbia Water and Power Authority
(BC Hydro) portfolio of dams and many of them are located in areas of high and moderate
seismicity. They range in age from under 20 years to over 80 years. It is assumed that a
probability of a failure of a concrete gravity dam is low, but such an event accompanied by
release of the reservoir would have extremely high consequences. Therefore, the problem of
dam safety against sliding is worth studying.
In an effort to enhance methods of seismic risk assessment of its concrete gravity dams, the B C
Hydro has initiated a parametric study on the seismic stability of low to medium height concrete

CHAPTER

INTRODUCTION

gravity dams. In an attempt to calibrate simple analytical methods used in risk assessment, B C
Hydro commenced in 1996 a phased experimental and analytical research project with U B C .
The studies described in this thesis were partially conducted within the B C Hydro - U B C
collaborative project. Some studies were conducted after the final phase of the collaborative
project was completed.

1.2 OBJECTIVES AND SCOPE OF THIS STUDY


The purpose of this study was to gain improved understanding of certain aspects of the dynamic
post-crack response of concrete gravity dams. In particular, the safety against base sliding of
existing concrete gravity dams was studied considering the non-linear effects at the damfoundation interface. A single monolith of a low to medium height concrete gravity dam was
considered with a through crack developed at the dam-foundation interface. It was decided to
address the problem by a series of studies including experiments, analytical work and reliability
analysis. The objectives and scope of the study can be grouped into three main areas:
1) The objective of the experimental work was to develop a scaled model of a single concrete
gravity dam monolith, and measure the amount of sliding of the unbonded model when this
was preloaded by a simulated hydrostatic force and subjected to different base excitations.
It was desired to investigate the response of the model for shake table induced base
accelerations of varied characteristics, amplitude and dominant frequency. The model was
unbonded at the base and only frictional characteristics of the dam-foundation interface were
simulated.

The characteristics of base excitations comprised harmonic and synthetic

earthquake records with dominant frequencies from 5 to 25 Hz.

CHAPTER

INTRODUCTION

During the experiments, the reservoir's loads were limited to simulated hydrostatic force.
Other effects, such as hydrodynamic force or uplift pressure at the foundation, were
neglected. It was not the objective of the study to directly apply the test results to predict the
seismic performance of a specific prototype.
2) The objective of the analytical study was to develop a simple numerical model to simulate
the sliding of a rigid block on a rigid foundation, preloaded by a constant horizontal force
and subjected to unidirectional horizontal base excitations. The requirement of the model's
simplicity came from the intended applications of the model for reliability analysis, which
typically involves a large number of simulations and, therefore, a fast analysis procedure was
preferable.

Another objective of this study was to develop a flexible model of the

experimental setup and use it to simulate some of the shake table tests conducted during the
experimental part of the study. This analysis was performed in order to find out how closely
the numerical models of varied complexity simulated the response of the experimental setup
measured during the shake table tests. The loads considered in the numerical models were
limited to those simulated during the shake table tests.
The objective of the analytical study of a full-scale monolith was to perform a series of
numerical simulations to calculate the response of the monolith with a varying water level
of the reservoir, characteristics of the earthquake record, and its peak ground acceleration
(PGA). It was the objective to perform these analyses using two different numerical models
and compare results. In addition to the forces simulated on the experimental model during
the shake table tests, hydrodynamic pressures were considered during this analysis. Other
loads, such as uplift or cohesion in the foundation interface were not simulated. Base
5

CHAPTER

INTRODUCTION

excitations on the monolith were limited to those in a single horizontal direction.


During the analysis, only the full-length crack between the dam monolith and the foundation
rock was considered. No weak horizontal construction joints in the monoliths were
considered and therefore, only the base sliding was calculated in the analyses. Should the
sliding stability be evaluated at other location than at the base-foundation interface, it would
have to be done in a separate study.
3) The objective of the reliability study was to formulate a procedure to obtain annual
probabilities of failure of the full-scale dam monolith using the results obtained from the
analytical study. Probability of failure was defined as a chance of exceeding a specified
amount of base sliding. The PGA, the error from the characteristics of the earthquake motion
and the reservoir's water level were considered as random parameters. Another objective
was to relate the results of the reliability study to safety factors against sliding.

1.3 THESIS OUTLINE


This section describes the manner in which the remainder of the manuscript is organised. The
thesis contains seven main chapters, each divided into several sections and subsections. In
addition, it includes five appendices with complementary information.
A study of available literature sources on experimental and analytical evaluation of sliding of
concrete gravity dams was essential for the project. Chapter 2 provides a comprehensive review
of published related research results in the last two decades.
Chapters 3 and 4 describe the experimental studies conducted during the project. The
6

CHAPTER

INTRODUCTION

preliminary experimental study, containing the description of the experimental model and setup,
is given in Chapter 3. Further testing including results from several types of experiments is
described in Chapter 4.
Chapters 5 and 6 deal with the analytical studies conducted within this project. Development,
verification and calibration of numerical models, and results of simulations of selected
experiments are described in Chapter 5. Chapter 6 contains several sections on the analytical
study of a full-scale concrete gravity dam monolith.
Chapter 7 describes several reliability studies utilizing the analytical data obtained in Chapter 6.
The chapter starts with description of basics of the reliability analysis, which is followed by
sections to identify random variables and presentation of results.
Chapter 8 provides a summary of the conducted research and a discussion of the conclusions
drawn. The need for further studies, both analytical and experimental, is outlined.

LITERATURE REVIEW

The project described in this thesis included three types of studies: experimental, analytical and
reliability. This chapter consists of a review of experimental and analytical studies in the area
of seismic response of concrete gravity dams. Focus is being placed on the experimental and
analytical evaluation of sliding of these dams. The reliability study was limited to using existing
methods and procedures and, therefore, no literature review for this study is given here. Instead,
fundamentals of the reliability analysis are summarised in the chapter describing this study.
The review of previous research conducted in the area of seismic behaviour of concrete gravity
dams was limited to the last two decades. The review begins with a section on performance of
concrete gravity dams during past earthquakes. This is followed by a section on experimental
studies investigating sliding of concrete gravity dams due to base excitations. Since such
response has never been measured on an actual dam, the experimental works referenced here
included mostly shake table testing of scaled models of single monoliths of concrete gravity
dams. The last three sections of this chapter contain a review of related research. In this review,
focus is being placed on reported research related to analysis of seismic-induced sliding of
concrete gravity dams.
It should be mentioned that two publications including a comprehensive literature review were
found very useful by the author when developing this literature review. One is the review by
Hall (1988), which presented a detailed summary of experimental work conducted on concrete
dams and their models before 1988. Analytical findings were included there i f numerical
simulations accompanied the experiments in a given study. The other review was performed by
Tinawi et al. (1998a) during a project of this research team in the area of structural safety of
8

CHAPTER

existing concrete dams (Tinawi et al., 1998b and 1998c).

LITERATURE REVIEW

This excellent summary of

experimental and analytical research addressed a wide range of topics involved in dam safety,
such as numerical modelling of lift joints, evaluation of flood and earthquake safety of gravity
dams considering lift joints and others.

2.1 CONCRETE DAMS DURING PAST EARTHQUAKES


There are several historical records of a concrete gravity dam undergoing shaking caused by an
earthquake. None of these cases ended with significant damage or collapse of the affected dam
and no base sliding of a dam was observed. The apparent good performance of concrete gravity
dams in past earthquakes should be viewed in the proper perspective. No dam on record, with
full reservoir, has been subjected to the most severe earthquake assessed possible for its
respective site (Fan and Sled, 1992).
The most significant reported event experienced by a concrete gravity dam was the 1967
shaking of the 103-m-high Koyna Dam in India with nearly full reservoir (Chopra and
Chakrabarti, 1972).

Recorded ground motions at the dam from a nearby earthquake of

magnitude 6.5 peaked at 0.49g in the stream direction and continued strongly for 4 sec. The
dam sustained with significant horizontal cracks through a number of nonoverflow monoliths at
an elevation of 36 m below the crest where the downstream face changed slope.
Perhaps the strongest shaking experienced by a concrete dam to date was that which acted on
the Lower Crystal Springs Dam in California, a curved gravity structure with modest height of
42 m, during the magnitude 8.3 San Francisco earthquake of 1906 (National Research Council,
1990). The dam incurred no damage even though it stood with its reservoir nearly full within
9

CHAPTER

LITERATURE REVIEW

350 m of the fault trace at a point where the slip reached 2.4 m. It should be mentioned that the
stability of this structure exceeds that of a typical gravity dam due to its curved plan and a cross
section that was designed thicker than normal in anticipation of future heightening, which was
never completed.
Another example of a concrete dam subjected to strong ground shaking is the 103-m-high
Pacoima Dam in Southern California during the 1971 San Fernando earthquake (Hall, 1988) and
during the 1994 Northridge earthquake (Stewart et al., 1994). This dam, and others like the
Ambiesta Dam in Italy during the 6.5 magnitude Friuli earthquake of 1976, the Rapel Dam in
Chile during the magnitude 7.8 Chilean earthquake of 1985, and the Gibraltar Dam in California
during the magnitude 6.3 Santa Barbara earthquake of 1925, represent cases of concrete arch
dams under severe earthquake shaking. However, analysis of such events is beyond the scope
of this research.
Another two examples of concrete dams subjected to severe base excitations are the 106 m tall
Sefi-Rud Dam in Iran hit by a 7.3 magnitude earthquake, which occurred near the dam
(Indermaur et al., 1992), and the 105 m high Hsingfengkiang Dam in China affected by a nearby
magnitude 6.1 earthquake (Chenet et al., 1974). Both these dams suffered some damage
including horizontal cracks at the level of abrupt change of the downstream slope and had to be
strenghtened and stabilised. Both these dams are buttress structures and the experience gained
from their performance during the earthquakes is not applicable to this study.
Although the experience outlined above representing the most significant earthquake events that
have acted on concrete dams, is impressive, it falls short of providing a complete confidence that

10

CHAPTER

LITERATURE

REVIEW

concrete gravity dams with full reservoirs are safe against strong seismic shaking.

2.2 EXPERIMENTS ON ACTUAL CONCRETE GRAVITY DAMS


A number of tests on prototypes of concrete gravity dams have been conducted in the last twenty
years. However, most of these were associated with low level vibrations, either ambient or
shaker-induced. Experimental research of non-linear effects in seismic response of actual
gravity dams due to strong ground shaking is very limited.
In the absence of prototype data, the need for validating numerical predictions with
experimental investigations has long been felt by the engineering community. However, due to
the difficulties involved in the modelling of dam-foundation-reservoir systems in the laboratory,
the experimental research has not kept pace with the numerical advances in the last two decades.
In the last decade, several shake table testing programs have been conducted with the objective
to study failure behaviour of scaled models of concrete gravity dam monoliths. Some of these
studies are summarised below.

2.3 EXPERIMENTS ON SCALED MODELS


In order to assess experimentally different types of non-linear seismic response of concrete
gravity dams, researchers conduct shake table tests on failure models. Most of the shaking table
tests in the past have been carried out using models of arch dams. In only a few cases have shake
table tests been conducted on models of concrete gravity dams.

11

CHAPTER

LITERATURE REVIEW

2.3.1 Niwa and Clough (1980)


Niwa and Clough (1980) described shake table tests carried out on a single 1:150 model
monolith of the 103 m high Koyna Dam. The model of the monolith was developed to maintain
similitude with the prototype. A plaster-based material used to build the model permitted the
simulation of response of the dam due to ground shaking in linear as well as nonlinear regions.
These tests demonstrated that a crack appeared in the model at the level of downstream slope
change, which was the location at which a crack appeared in the prototype structure during the
1967 Koyna earthquake.

2.3.2 Donlon and Hall (1991)


Another series of shake table tests was reported by Donlon and Hall (1991). These tests were
conducted on a 1:115 length-scaled model of the 122 m high single monolith of the Pine Flat
Dam. A total of three models were tested, out of which only one was made of a single
monolithic plaster based material. The other two models were constructed with composite
materials consisting of a weak upper half portion, made of a plaster-based material, and a
stronger main body made of a polymer-based material. The tests showed a crack initiation just
below the neck of the dam, although owing to the differences in the material properties and the
construction, the crack profile in the three models was not consistent and differed significantly
from one another.

2.3.3 Zadnik and Paskalov (1992)


A series of shake table tests on a concrete model of a single gravity dam monolith was conducted
by Zadnik and Paskalov (1992). This testing quantified sliding of the monolith, resting on
12

CHAPTER

LITERATURE REVIEW

either wooden, concrete or steel floor, due to ground motions generated by the shaking table.
Results of these tests were compared with numerical predictions with a good agreement. The
experimental model used during these tests was not developed to satisfy similitude with any
actual dam.

2.3.4 Lin etal. (1993)


Lin et al. (1993) performed a series of shake table tests on five 1:130 scale homogenous models
of a concrete gravity dam 195 m high. The models were excited harmonically in the first mode
of vibration until the first crack occurred near the neck of the dam. Other cracks appeared later
in the middle part of the model. Finally, cracks at the toe and the heel became visible. However,
no dam-reservoir interaction was modelled during the tests. If this was taken into consideration,
as mentioned in the study, the crack at the heel was likely to appear much sooner.

2.3.5 Mir and Taylor (1995)


A series of shake table testing of a gravity dam monolith model was conducted by Mir and
Taylor (1995). These tests were conducted to identify crack pattern of a 1:30 scaled model of
30 m high gravity dam.

The model has been developed from plaster based material in

accordance with similitude laws. The model was fixed securely to the shake table platform,
therefore, foundation flexibility was not simulated during these tests. Reservoir effects were
simulated by hydrostatic and hydrodynamic (Westergaard added mass approximation)
pressures. A total of eight models were tested under different three conditions: no reservoir
effects, only hydrostatic pressure, both hydrostatic and hydrodynamic pressures. The models
were excited by harmonic ground motion at 60 Hz frequency as well as by earthquake records.

13

CHAPTER

LITERATURE

REVIEW

The amplitudes of shaking were increased until failure occurred. Base cracking was observed
to be the main failure mechanism and tendency of the models to slide and rock after the full
crack development at the interface was also observed in some cases.

2.3.6 Mir and Taylor (1996)


A series of shake table tests complementary to the previous study (Mir and Taylor, 1995) was
conducted by M i r and Taylor (1996). Several 1:30 models of a 30 m gravity dam were
constructed from concrete since the strength of the model was not the controlling parameter.
The interface between the dam and its foundation had only frictional resistance with the static
coefficient of friction of 0.72. Under full reservoir condition and with no uplift pressure
considered, downstream sliding was identified as the only significant global failure mechanism.
For an empty reservoir condition, it was observed that the dam model might not slide, but
instead it could overturn upstream about its heel for high interface frictional resistance.

2.3.7 Tinawi etal. (1998 b, c)


A n extensive shake table experimental program to study seismic sliding response of three 3.4 m
high concrete dam models with a weak lift joint at mid-height was carried out by Tinawi et al.
(1998b). The hydrostatic and hydrodynamic forces were approximated with a hanging weight
attached to the specimens. The models were subjected to various base excitations including
single and double triangular pulses and earthquake accelerograms of varying P G A and different
frequency content. It was observed from the measured cumulative sliding displacements that
sliding of the specimen due to a single pulse of sufficient duration can be significant. Further,
it was concluded that the sliding due to Western North American earthquake records was about

14

CHAPTER

LITERATURE REVIEW

three times larger compared to Eastern records for a given PGA. Other conclusions, such as the
amount of sliding was proportional to the P G A raised to a power of about three, were also made.
Most of the experimental studies mentioned in this section were conducted together with
numerical simulations of the tests. Some analytical methods to study seismic response of
concrete gravity dams will be summarised in the next sections.

2.4 TRADITIONAL SEISMIC ANALYSIS OF CONCRETE


GRAVITY DAMS
Concrete gravity dams were traditionally designed for static loads up to about mid 1960's. The
earthquake forces were treated simply as static equivalent forces and were combined with the
hydrostatic pressures and gravity loads. The analysis was concerned with overturning and
sliding stability of the monolith treated as a rigid body and with stresses in the monolith, which
were calculated by elementary beam theory. In representing the effects of the horizontal
upstream/downstream ground motion by static equivalent lateral forces, neither the dynamic
response characteristics of the dam-water-foundation system nor the characteristics of the
earthquake ground motion were recognized. Two types of static lateral forces were included:

Forces associated with the weight of the dam were expressed as the product of a
seismic coefficient, typically constant over the height with a value between 0.05
to 0.1, and the weight of the portion being considered.

Hydrodynamic pressures, in addition to the hydrostatic pressure, were specified


in terms of the seismic coefficient and a pressure coefficient, which was based on
assumptions of a rigid dam and incompressible water.

15

CHAPTER

LITERATURE REVIEW

Interaction between the dam and the foundation rock was not considered in computing the
aforementioned earthquake forces.
The traditional design criteria required that an ample factor of safety be provided against
overturning, sliding and overstressing under all loading conditions. Tension was often not
permitted, even if it was, the possibility of cracking of concrete was not seriously considered. It
had generally been believed that stresses were not a controlling factor in the design of concrete
gravity dams, so that the traditional design procedures were concerned mostly with satisfying
the criteria for overturning and sliding stability. It was apparent from observed behaviour of
dams subjected to ground shaking, such as Koyna Dam, that stresses thus calculated in gravity
dams due to standard design loads have little resemblance to the stresses due to the dynamic
response of such dams to earthquake ground motion.
The discrepancies between calculated and observed behaviour of the dams were partly the result
of not recognizing the dynamic response due to earthquake motions in computing the
earthquake forces included in the traditional design methods. The typical values used for the
seismic coefficient, ranging from 0.05 to 0.1, were much smaller compared to ordinates of
pseudoacceleration response spectra for intense earthquake motions in range of vibration
periods up to 1 sec. In addition, traditional analysis and design adopted a uniform distribution
of the seismic coefficient, ignoring the vibration properties of the dam. This kind of seismic
coefficient lead to underestimating stresses in the upper part of dams and it ignored dynamic
influence of the widening of the dam cross-section for non-structural reasons near the dam crest.
Another source of discrepancies between calculated and observed behaviour of the dam was not

16

CHAPTER

LITERATURE

REVIEW

recognizing hydrodynamic effects, which were found to be grossly underestimated in the


traditional design loading. When the compressibility of water and dam-water interaction
resulting from deformations of the dam are included in the analysis, hydrodynamic effects are
generally shown to be important in the response of concrete gravity dams. For example, in the
analysis of the response of nonoverflow monolith of the Pine Flat Dam, on rigid rock
foundation, due to the Kern County, 1952 earthquake (Taft Lincoln School Tunnel record), the
tensile stresses in the dam were approximately 30% larger when the hydrodynamic effects were
included compared to the case when they were not (Fenves and Chopra, 1985) included.
Foundation rock flexibility is another phenomenon, which was not considered in computing
earthquake forces in traditional design loading. When dam-foundation rock interaction is
properly included in the dynamic analysis, these effects are generally significant and they
mostly reduce stresses in the dam (Fenves and Chopra, 1985).
The above list of significant phenomena which were not included in the traditional analysis of
concrete gravity dams stimulated the development of more advanced tools to analyse response
to earthquake loading.

2.5 REFINED SEISMIC ANALYSIS OF CONCRETE GRAVITY DAMS


a

Realistic analyses of the seismic response were not possible until the development of the finite
element method, recent advances in dynamic analysis procedures, and the availability of largecapacity, high-speed computers.

Thus, much of the research involving linear time history

analyses did not start until the mid-1960's.

17

CHAPTER

LITERATURE REVIEW

Initially, all nonlinear effects, including those associated with construction-joint opening,
concrete cracking and water cavitation, were ignored, and the interaction effects were either
neglected or grossly simplified. Subsequently, development of special techniques permitted the
incorporation of the interaction effects into linear analyses. These effects included phenomena
such as dam-water interaction based on dam flexibility and water compressibility, radiation
damping into reservoir bottom, and dam-foundation rock interaction (Fenves and Chopra,
1985a). Based on these techniques, numerical models implemented in computer programs have
been developed. Most of the models operated in two dimensions although three-dimensional
models have also been developed (Fok et al. 1986).

The two-dimensional analysis was

recommended whenever it was appropriate for the dam to be analysed, because it is


computationally efficient and it could rigorously consider most of the factors significant for the
earthquake response. In the three-dimensional analysis, dam-foundation rock interaction and
earthquake excitation were usually treated in an overly simplified manner.
Nonlinear phenomena in concrete gravity dam seismic analysis began to be addressed in the
1980's. Nonlinear analysis presents a number of problems, such as the modelling of input
motions, modelling of horizontal construction joints in the monoliths, nonlinear material
behaviour of concrete and rock, and uplift of the dam. Contrary to rigorous elastic solutions,
which mostly operate in frequency domain, a nonlinear analysis requires a time stepping
solution, and incorporating infinite boundaries in such models is still not reasonably well
understood (Taylor, 1996). As most dam engineers and researchers have access to commercial
finite element packages such as A B A Q U S , ADINA, A N S Y S and others, which have powerful
nonlinear capabilities, methodologies to exploit these tools were developed together with the

18

CHAPTER

LITERATURE REVIEW

development of the specialized codes.


One of the most important nonlinearities is cracking of concrete. Cracking can develop, for
example in the regions of pre-existing thermal cracks or due to overstressing of concrete during
strong shaking by an earthquake. Various post-yield constitutive relationships have been put
forward for dam concrete. Cracking was first investigated with smeared crack assumptions and
also with plasticity models, and later with the application of fracture mechanics (Reich,
Cervenka and Saouma, 1991). Simic and Taylor (1995) compared a smeared cracking model
with a plasticity model applied to a gravity dam, which was subjected to a range of hydrostatic
loads. This comparison showed how the two models tend to predict very different failure
modes. The plasticity model lead to widely distributed yielding, while the smeared cracking
model predicted localized cracking that was more consistent with crack propagation through a
brittle material. Simic and Taylor (1996) also discussed the application of a typical smeared
crack model available in codes such as A D I N A and SOLVIA.
Possible cracking of low to medium height gravity dams would occur, in contrast to tall dams
such as Koyna or Pine Flat Dams, at the dam-foundation interface. This was shown by nonlinear analysis for the 56-m-high Russel Dam by Mlakar (1987). This analysis yielded extensive
cracking along the base of the dam under strong ground shaking, which was different from the
crack pattern in the taller dams considered in the same study. Therefore, base sliding of a dam
with a crack through its base has been studied by several researchers in recent years.
Several types of nonlinearities could be studied in the seismic response of concrete gravity
dams. The study described in this thesis focused on one of them, the base sliding response of

19

CHAPTER

concrete gravity dams during an earthquake.

LITERATURE

REVIEW

Only the references related to this type of

nonlinearity are summarised in the next section.

2.6 SEISMIC ANALYSIS RECOGNISING EXPLICITLY


THE DAM-FOUNDATION CONTACT PLANE
Sliding response of concrete gravity dams, defined as a concentrated shear deformation at the
interface between the dam and the foundation rock, has been studied in the last decade. This
was due to earlier analytical and experimental research, which had identified that the crack
pattern in low to medium dams under an earthquake might lead to development of a crack
between the dam and its base.

2.6.1 Leger and Katsouli (1989)


A study of the seismic stability of a 90 m concrete gravity dam was reported by Leger and
Katsouli (1989). Gap friction node-to-node finite elements were used to model the damfoundation interface. The dam and the foundation were assumed to remain elastic. The contact
elements were assumed to have tangential and normal compressive strength but no tensile
strength. Thus, the sliding and rocking response of the dam monolith could be simulated. It was
found that the maximum base shear values computed at the dam-foundation interface were
greater for linear analyses than for nonlinear, especially for sites with flexible foundation
conditions, as compared to the dam concrete. A dynamic stability safety factor was defined as
a ratio of the specified design P G A to the critical acceleration required to reach a critical safety
limit defined as significant sliding displacement or loss of bond over more than 90% of the base.

20

CHAPTER

LITERATURE REVIEW

2.6.2 Chopra and Zhang (1991)


Simplified analytical formulations to evaluate earthquake-induced sliding of a concrete gravity
dam monolith were derived by Chopra and Zhang (1991). The dam monolith was assumed to
be supported on a plane surface without bonding. Different formulations were developed for
rigid and flexible dams considering only the contribution of the fundamental mode of vibrations.
Downstream sliding was found to be the only significant response of the dam. In addition, it
was found that the dam flexibility had the effect of increasing the permanent sliding
displacement as compared to the rigid case. Vertical ground motions were found to increase the
total amount of sliding during an earthquake.

2.6.3 Danay and Adeghe (1993)


Empirical formulas to obtain sliding displacement of a concrete gravity dam were derived by
Danay and Adeghe (1993). These formulas were derived for the dams of 60 m or less subjected
to Eastern North America as well as Western (California) type of earthquake ground motions.
The empirical formulas were obtained from statistical regression on the results of a series of
analyses performed with a sliding block model with the dam-foundation interface represented
by gap-friction elements. The slip induced by horizontal acceleration alone was found to be
approximately 60% of that obtained when horizontal and vertical accelerations were applied
simultaneously.

2.6.4 Chavez and Fenves (1995 and 1996)


The work presented by Chavez and Fenves (1995) included dynamic analysis of sliding of an
earthquake-loaded gravity dam and compared the results of this analysis with those from the
21

CHAPTER

LITERATURE REVIEW

traditional check of sliding stability. This involved computing a factor of safety against sliding,
calculated as a function of a Mohr-Coulomb model of concrete and foundation material friction
characteristics, self weight of the dam, hydrostatic pressure, uplift force, cohesion force at the
dam-foundation interface zone, and equivalent static loads on the dam representing the dynamic
effects of an earthquake.

Figure 2.1: Model of the Dam-Water-Foundation System Used in Program EAGD-SLIDE

The work of Chavez and Fenves (1996) applied a novel modelling and solution algorithm, the
hybrid frequency-time domain procedure (Kawamoto, 1983; Darbre and Wolf, 1988). The
numerical model developed during this work accounted for the dynamic characteristics of the
dam, foundation rock flexibility, compressible water, and Mohr-Coulomb model for base
sliding. This numerical model was implemented in computer program EAGD-SLIDE (Chavez

22

CHAPTER

LITERATURE REVIEW

and Fenves, 1994). The idealized dam-foundation-water system used in program E A G D SLIDE is shown in Fig. 2.1. Using the program, the dam might slide along the rigid interface
between the dam base and the foundation rock. Rocking of the dam about the base is not
represented in the model. The system is subjected to horizontal and vertical components of freefield earthquake ground motions at the base of the dam.
The body of the dam in EAGD-SLIDE is modelled by two-dimensional plane stress finite
elements with linear elastic properties. The water impounded in the reservoir is idealized as a
two-dimensional domain extending to infinity in the upstream direction. The water is treated as
an compressible fluid that produces hydrodynamic pressures depending on the excitation
frequency. The foundation rock is idealized as a homogeneous, isotropic and viscoelastic halfplane. The interface between the dam monolith and the foundation rock (foundation interface)
is assumed to be straight surface with the resistance to sliding governed by Mohr-Coulomb law.
The resistance at the interface depends on a cohesion (bonding) force, the coefficient of friction,
and the time varying normal force. The model assumes that sliding occurs along the entire
interface.
The work of Chavez and Fenves (1996) included a parametric study on a 122-m-high monolith
of the Pine Flat Dam. Several combinations of input parameters of dam-foundation-water
system were considered. Depending on these parameters, base sliding from 0 to 29 cm was
obtained for three different historical earthquake records with P G A of 0.4g. The study identified
that the foremost factor influencing the amount of sliding is foundation rock flexibility; rigid
foundation rock yielded much larger sliding than flexible foundation rock. The sliding was also
sensitive to the coefficient of friction and cohesion of the interface zone. Based on their
23

CHAPTER

LITERATURE REVIEW

conclusions, the authors proposed a framework for evaluation sliding stability of gravity dams.

2.6.5 Mir and Taylor (1996)


A series of finite element analyses were performed together with the shake table tests by Mir
and Taylor (1996). The analyses using a nonlinear large displacement contact surface algorithm
were carried out in order to establish comparison between the measured and calculated sliding
displacement responses.

Generally, a good agreement was obtained, even though some

discrepancies were observed for the magnitude of displacements under some of the peak pulses.
In these cases, it was observed that numerical analyses underestimated sliding displacements
under earthquake loads compared to the measured values. Conclusions suggested that the actual
interface dynamic friction coefficient was transient in nature and lower than the friction
coefficient.

Additional analytical work was carried out to assess the validity of simplified

sliding analysis methods for dams considered as rigid bodies.

2.6.6 Tinawi et al. (1998b, c)


A series of numerical studies were included in the research reported by Tinawi et al. (1998b, c).
During these studies, several lift joints constitutive models were developed using thin layer
interface finite elements and gap-friction elements. These were implemented into computer
programs to perform numerical simulations of the flood and earthquake responses of the three
gravity dams (17.9,90 and 116 m high).
The finite element seismic response analysis of a 90 m gravity dam model including 8 lift joints
along the height lead to several conclusions. The presence of weak joints along the dam's height
reduced the shear force and associated residual displacements acting at the dam-foundation
24

CHAPTER

LITERATURE REVIEW

interface, as compared with the case where only the base joint was modelled. However,
reductions in base sliding displacements were associated with significant relative sliding
responses of upper lift joints.
It was found that the presence of weak lift joints along the height of a gravity dam can affect the
peak ground acceleration to induce sliding. A n upper bound estimate of base displacements can
be obtained by considering a model with a single joint at the base. However, to ensure the
stability of the upper section of the dam, the presence of weak lift joints along the height should
be explicitly considered in safety evaluation.
The magnitude and spatial distribution of residual sliding displacements were found dependent
on the frictional strength, and therefore on the surface preparation and subsequent aging of the
joints.

A n upper bound estimate of residual sliding displacements could be obtained by

considering that joints have no tensile or cohesive strength. Static values of friction coefficient,
which controls the initiation of sliding, should be selected conservatively i f the analysis is
performed to assess the magnitude of residual sliding displacements. That is because for a given
magnitude of compressive force, the dynamic frictional strength is likely to be smaller than its
corresponding static value.
In addition, the analyses included correlations between numerical simulations and experimental
seismic sliding responses of the 3.4 m tall dam models tested previously. Good correlations with
experimental results were obtained using two numerical models, one based on rigid body
dynamics and frictional strength limited by the Mohr-Coulomb criterion, and the other based on
finite elements using gap-friction elements.

25

3 PRELIMINARY EXPERIMENTAL
STUDY OF A DAM MODEL
Shake table tests are increasingly being used to evaluate seismic performance of structures, their
models or various structural and nonstructural components. Shake table testing is a valuable
tool especially in cases when the response being studied is associated with failure, such as in the
sliding of a model of a concrete gravity dam. In such cases, this type of testing is the only source
of experimental data to calibrate numerical models.
In December 1996, a series of shake table tests on a scaled model of one monolith of a concrete
gravity dam was conducted at the Earthquake Engineering and Structural Dynamic Research
Laboratory (Earthquake Laboratory) at the Department of Civil Engineering of the University
of British Columbia (UBC). The tests were conducted as the first phase of a collaborative
research between U B C and the Dam Safety Program at the Maintenance, Engineering and
Projects Division of the British Columbia Hydro and Power Authority (BC Hydro). The goal
of this research was to gain an improved understanding on the post-crack dynamic response of
concrete gravity dams. These tests were also the first of two series of experiments, which
yielded data needed for this study. In this thesis, this experimental program is referred to as
preliminary tests.

3.1 OBJECTIVES AND SCOPE OF PRELIMINARY TESTS


The objective of the preliminary tests was to measure how much a model of a single monolith
of a gravity dam would slide if preloaded by a simulated hydrostatic force and subjected to base
excitations. The work focused on frictional characteristics of the dam-foundation interface and

26

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

therefore, the model was unbonded at the base during the tests. Other forces at the interface,
such as cohesion, were not simulated. The excitations were limited to harmonic motions only.
In addition, there were several other objectives of the tests:

To assess the feasibility of using U B C s Earthquake Laboratory to conduct this


type of tests.

To develop sufficient experimental data for the calibration of computer models to


simulate sliding of concrete gravity dams unbonded at the base.

To tune-up the experimental setup before a more extensive series of the shake
table tests, which were already being planned at the time of the preliminary tests
as a second phase of the collaborative U B C - B C Hydro project.

The forces acting on the model were limited only to earthquake motions simulated by the shake
table and to hydrostatic pressure represented by a single horizontal force on the model. Other
effects, such as hydrodynamic force or uplift pressure at the foundation, were neglected. It was
not the objective of the study to directly apply the test results to predict the seismic performance
of a specific prototype.

3.2 DESCRIPTION OF TESTING FACILITY


The Earthquake Laboratory at U B C laboratory is 16.4 m long and 11.5 m wide, providing space
for construction, assembly, and handling of relatively large structural models and heavy
equipment. Access is available for direct entry of large vehicles.
The laboratory is equipped with an advanced, closed-loop, servo-controlled hydraulic seismic

27

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

simulator or shake table. It can accurately reproduce earthquake ground motions in one or more
directions. The main element of the table is a 3 m by 3 m platform, which consists of a 0.4 m
thick aluminium cellular structure. The weight of the table is about 20.5 k N and it has a grid of
38 mm diameter holes that are used to attach test specimens. The aluminium platform and
attached hardware were designed to have a fundamental frequency about 40 Hz, so that it can
be considered rigid within the operating frequency range of the shake table tests, which is mostly
1 to 25 Hz. Clearance above the table is 4.2 m and the laboratory is equipped with a 44.5 k N
overhead crane for placing models and equipment on the shake table. The shake table has a
pay load capacity of 156 k N and it can be configured to produce two types of multi-directional
motions. One configuration (called

1HX3V

- 1 Horizontal and 3 Vertical) can be used to

simulate longitudinal, vertical, pitch and roll motions. The other configuration (called 3H) can
be used to simulate longitudinal, lateral, and torsional (yaw) motions. In configuration 3H,
horizontal, longitudinal motions are produced by one hydraulic actuator with a maximum peakto-peak displacement of 15.2 cm (6 inches). This actuator has a main stage area of 80.9 cm .
2

Other two actuators, with 45.2 c m of effective piston area each, could produce horizontal
2

motions in the transverse direction.


The actuator force reactions are resisted by a massive pit. The pit is a reinforced concrete
foundation extending around the table in the form of an open box with the thickness of 1.5 m on
the sides where the actuators are installed. The outside plan dimensions of the wall are 6 m by
5.5 m and it is 2.5 m high, while the inside dimensions are 3.6 m by 3.6 m by 2 m (Rezai, 1999).
The displacement of the table is limited by the stroke of the actuators, 7.6 cm. The flow rate
in the servovalves limits the maximum velocities produced in both directions to 100 cm/sec.
28

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

The maximum acceleration is limited by the force limits of the actuators together with the mass
of the table-specimen system. The stalling force capacity of the longitudinal actuator is 156
kN, while the other two actuators in the transverse direction have a capacity of 90 k N each.
The shake table is controlled by a signal processing subsystem, driven by a replication multishaker control software. This software performs a closed-loop control of several shakers which
are capable of replicating recorded earthquake shaking and other types of motions with high
accuracy. The high performance digital control system can easily replicate earthquake motions
for models with different mass-stiffness characteristics.

This is a desirable feature for

comparative studies of different models or equipment under the same loading conditions.
The data acquisition system at the laboratory can record up to 128 channels of instrumentation
information from a test specimen. From this number, a total of up to 44 channels can be
conditioned by variable gain buffers and cut-off filters, which provide optimal control over
signal levels and noise reduction in order to retrieve accurate dynamic testing data.

3.3 EXPERIMENTAL MODEL


A n experimental model for a single monolith of a concrete gravity dam is described in this
section. The model consisted of three parts: a dam monolith, a pair of surfaces plates and
hydrostatic load simulation assembly. Description of each part is given here in a separate
subsection and the model without the hydrostatic load assembly is shown in Figure 3.1

3.3.1 Shape of the Model for a Dam Monolith


The model for a dam monolith was 1500 mm high, 1250 mm wide in the upstream/downstream

29

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

Figure 3.1: Main Parts of Experimental Model


direction and 480 mm thick in the cross canyon direction. The shape of the monolith, shown in
Figure 3.2 was designed close to a shape, which could have a monolith of a low to of medium
height concrete gravity dam. The model was designed with several holes for easier handling.
In an optimum case, the experimental model should be of the same dimensions and made from
the same material as a prototype. This is mostly not possible for models of large structures due
to given dimensions of labs and limited payloads of earthquake simulators. In addition to this,
during this testing the model was handled very often because it had to be put back to its original
position after each test when it moved.

30

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

225

1250

00

1250
Figure 3.2: Schematic of the Model for a Dam Monolith

3.3.2 Material for Monolith and Similitude Considerations


It was considered during the design of the model that it would be moved often back to its original
position during the testing. Although the Earthquake Laboratory at U B C is equipped with a
crane it would be difficult to handle a heavy model of the monolith. It was decided to keep its
weight below 5 kN. The material for the model was designed based on this requirement.

31

CHAPTER

PRELIMINARY

EXPERIMENTAL

STUDY

OF A DAM

MODEL

The material used for the monolith was a mix consisting of Portland cement type 10, perlite,
styrofoam, silica fume and water. The weight composition of the mix was 42.2% cement, 40%
water, 12.5% perlite, 3.6% silica fume and 1.7% Styrofoam (Horyna et al., 1997). Since the
strength of the above material was low it was decided to use stronger material at the locations
of expected local loads on the model. These locations included about 5 cm thick layer at the
bottom of the model, which was in contact with the upper surface plate and also the part of the
model where the simulated hydrostatic force was applied. A modified mix without Styrofoam
was used for these parts of the model. The total weight of the model including the upper sliding
plate and all attachments was 480.6 kg.
It was recognised that the test results would be used to study behaviour of the model, but they
would not be used to predict response of any prototype dam. However, once the model was
built, it was of interest to find the dimensions of a full scale concrete gravity dam monolith
which would satisfy similitude with the model. Discussion of the similitude aspects follows.
To maintain similitude between inertia and elastic forces requires the quantities of length (Z),
mass density ( p), modulus of elasticity (E) and acceleration due to gravity (g) of a prototype
satisfy certain relationships with those of the experimental model (Moncarz an Krawinkler,
1981, or Mir and Taylor, 1995). Based on these quantities the following ratios can be defined:

PM

SM

where indeces P and M stand for prototype and model, respectively. These scaling factors
should satisfy the following condition resulting from Cauchy's requirements for proper
simulation of inertial forces and restoring forces:
32

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

(3.2)

The material characteristics of the prototype dam monolith made of concrete were considered
E = 27,000 MPa and P
P

= 2,580 kg/m (Powertech Labs, Inc., 1996). The material for the
3

model had the following characteristics: E = 500 MPa, which was obtained as a modulus of
M

elasticity measured on a testing cylinder; and o

= 720 kg/m , which was found by weighting

the testing cylinder. These values were related only to the basic material of the monolith, but
they did not take into account neither rebars in the model nor the parts of the model which were
made using stiffer and heavier mix without Styrofoam in it. Therefore, the values taking into
account all materials in the model should be used for similitude calculations. Such values were
found as: E = 560 MPa and p
M

- 800 kg/m . The new value for is^was obtained in the


3

analytical part of this study as the modulus of elasticity, which resulted in the best match
between natural frequencies of the model obtained experimentally with those from F E study.
The new value for p

was obtained after weighing the complete model. It can be concluded

that these new values for the characteristics of the model's material are higher than the original
one obtained from the testing cylinder. This increase of material properties against those of the
testing cylinder is reasonable and can be explained by the fact that the material used in some
parts of the model was stiffer and heavier than that of the testing cylinder.
Knowing the prototype's and model's material characteristics the ratios 5^=3.22 and S = 48
E

were calculated. Then, using the Cauchy's condition and considering S = 1, the scaling factor
for length can be calculated as S = 15. This value indicates that i f the dimensions of the
L

prototype are 15 times as large as those of the model, the similitude between elastic and inertia

33

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

forces would be satisfied. Since the height of the model is 1.5 m such a prototype would be a
concrete monolith about 22.5 m tall which is a height of a low concrete gravity dam. Once
knowing the length ratio S , the ratio of prototype to model frequencies S^-can be determined as
L

(Mir and Taylor, 1995):

(3.3)

The frequency ratio for the model considered can be evaluated as Sf= 0.258, which means that,
for example, the frequency of 10 Hz on model scale corresponds to 2.58 Hz for the prototype.
The similitude between elastic and inertia forces was discussed in this subsection. Although it
was not the objective of the experimental study to simulate response of any specific prototype,
this discussion was presented here in order to relate the model to an actual structure.

3.3.3 Sliding Surfaces


The model comprised the monolith and a pair of sliding surface plates, which were identified as
the Upper Surface Plate (USP) and the Lower Surface Plate (LSP). The experimental model was
designed in such a way that sliding occurred between the sliding surface attached to the bottom
of the dam and one attached to the shake table. The arrangement of model components provided
a convenient and easy way to change pairs of plates. In addition, it was required that the sliding
surfaces had stable friction characteristics during many tests.

The sliding surfaces were

designed according to the following requirements.


The dimensions of USP and LSP are given in Figures 3.3 and 3.4, respectively. The USP was
clamped to the bottom of the model as can be seen in Figure 3.5. The LSP was bolted to the

34

CHAPTER

PRELIMINARY

EXPERIMENTAL

STUDY

OF A DAM

MODEL

1150

Plan View
1150
o
m

Welded "Steel mesh


wire 3mm, squares 6'
Note:
all dimensions are in mm,
unless otherwise specified

Section A-A
480

Section B-B

o
in

Figure 3.3: Dimensions of the Upper Surface Plate


shake table using a steel frame, which is shown in Figure 3.6.
A total of three pairs of sliding surfaces were used for testing. These were denoted as:

Smooth surface;

Cracked-ice surface;

Rough surface.

It should be noted that these names were given to the surfaces during the fabrication process and
independent of how each surface performed during the testing. The plates for sliding surfaces
were made of a mortar based mix with fine aggregates only; 100 kg of this mix comprised 10.8

35

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM

MODEL

1500

Plan View
660

Welded SteefMesh,
wire 3 mm, squares 6

Section C-C
1500

o
LO
z x

Section D-D
Note: all dimensions are in mm, unless otherwise specified.

Figure 3.4: Dimensions of the Lower Surface Plate


kg of water, 24.5 kg of Portland cement type 10, and 64.7 kg of fine-grain play sand. The
fabrication process of the sliding surfaces for different roughnesses is explained below.
3.3.3.1

Smooth Surface

The Smooth surface was formed by casting first the Lower Surface Plate (LSP) placed on a
plastic sheet (1.5 mm thick, with smooth surface) at the bottom of the formwork. After curing,
the LSP and the plastic sheet were turned upside down and then the Upper Surface Plate (USP)
was cast on top. The same 1.5 mm thick plastic sheet was used to separate the two plates.

36

CHAPTER

300

PRELIMINARY EXPERIMENTAL

600

STUDY OF A DAM MODEL

300

1200
b) Plan View

Notes: all dimensions are in m m , unless otherwise specified.


(1)
(2)
(3)
(4)
(5)
(6)
(7)

steel plate 1/2", size 489*140 mm with holes for bolts(4)


rebars 5 mm, welded to plate (7)
rebars 10 mm, welded to plate (7)
bolts M14
nuts welded to (1)
spacer - steel plate 1/8" or 1/4", 480*35 mm
steel plate 1/4", 480*300 mm

Figure 3.5: Connection of USP to the Monolith

3.3.3.2 Cracked-ice surface


The Cracked-ice surface was formed by casting the LSP on a plastic sheet (1.5 mm thick, with
a Cracked-ice pattern surface). After curing, the LSP was turned upside down and the USP was
cast on top. A very thin plastic foil (wrapping foil with thickness of hundredths of mm) was

37

CHAPTER

PRELIMINARY

EXPERIMENTAL

STUDY

OF A DAM

MODEL

1720
315

200

340

330

*4 200

200

, 135

415

530

540

o
o

CM

L 100x75x10-1720 mm
1520
L 100x75x10 - 680 mm

o
00

oo
60 u

L100x75x10,
680 mm /

o
CO

L 100x75x10 - 1720 mm

o
o

o
o

CM

Hole, ID 40 mm (Typical)
a) Plan View
3/4" Bolt and Welded Nut

Plate 1/4"
200x200 mm (Typical)
Note: all dimensions are in mm,
unless otherwise specified.

rop of the Shake Table


b) Section F-F
Figure 3.6: Steel Frame for the LSP to Shake Table Connection
used as separator between LSP-top surface and USP-bottom surface.

3.3.3.3

Rough surface

The Rough surface was formed by casting the LSP on a plastic sheet (1.5 mm thick, with smooth
surface). After curing, the LSP was turned upside down and the top surface of the LSP was
chipped with a chisel. This produced randomly positioned irregularities (about 8 at an area 100

38

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

mm by 100 mm) with depths of 4 to 8 mm. This irregular surface was used as a bottom of the
formwork to cast the USP. Thin wrapping plastic foil was used as a separator between the LSPtop surface and the USP-bottom surface.

3.3.4 Simulation of Hydrostatic Load


It was the objective of the preliminary test to investigate response of the model due to base
excitations and simulated hydrostatic force. This force was provided by a hydrostatic load
simulator, which pulled the model in the downstream direction. The force was simulated by a
set of long springs stretched between the downstream side of the model and the support fixed to
the shake table. The force in the springs was kept relatively constant during the test as the model
moved. This was achieved by measuring the force during the test with a load cell and by using
a manual valve control to adjust the spring tension as needed. This can be seen in a photo in
Figure 3.7.

3.4 EXPERIMENTAL SETUP AND TESTING PROCEDURES


The testing program consisted of two types of tests, which were called static and dynamic. Both
types of the tests were conducted using the shake table. The objectives of the static tests were
to determine friction coefficients of the surfaces as well as the magnitude of static forces
necessary to cause sliding of the model preloaded by a specified hydrostatic force. The dynamic
tests were conducted to find the amounts of model sliding for varied frequency and amplitude
of base acceleration. This was desired to find for three different surfaces (Smooth, Cracked-ice
and Rough) and two levels of simulated hydrostatic force (70% and 100% of full reservoir).

39

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

Figure 3.7: Controlling the Level of Hydrostatic Force during a Dynamic Test
The 3 H configuration of the shake table was used to conduct all the tests reported here.
However, during all the tests conducted within this study, only longitudinal horizontal shake
table motions were used. This direction corresponded to the upstream/downstream direction of
the model for the dam monolith.

3.4.1 Static Tests


The experimental setup for the static tests is shown schematically in Figure 3.8 or in a photo in
Figure 3.9. It comprises the dam monolith, a pair of sliding surfaces, a set of springs and a rigid
bar. The latter was used to keep the model at the same position when the shake table was
moving slowly in the upstream direction (relatively with respect to the model). The springs
were used to apply previously computed force, representing the hydrostatic pressure on the
model. This force was kept relatively constant during the test as the model moved. During the
40

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

D a m Monolith,
Model"

-Support for Springs


Springs t o Simulate
J
Hydrostatic Force

rt

L o a d Cell 2
-Rigid B a r

VWWA

i
Pinned /
Support

L o a d Cell 1

S h a k e Table

Direction of S h a k e Table
M o t i o n During Static Tests

\ Vertical S u p p o r t
" (Hoist)
Concrete
Wearing
Surfaces

Figure 3.8: Experimental Setup for Static Tests

Figure 3.9: Photo of Experimental Setup for Static Tests

41

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

static tests, the rigid bar pushed against the model through a steel plate with a hinge connection.
The force between the model and the rigid bar was measured using another load cell.
A typical static shake table test consisted of the following steps:
Set up the model on the shake table.

Attach horizontal long springs to downstream face of the model at specified height
and pre-stretch them sufficiently to provide a force equivalent to the hydrostatic
force due to the reservoir.

Conduct the test by holding the model with the rigid bar attached to its upstream
face and slowly moving the table upstream at a constant velocity. Measure the
model displacement relative to the shake table and the force between the model
and the rigid bar. Maintain hydrostatic force at constant level during the test.

Detach the horizontal springs and lift the model.

Clean the sliding surfaces.

3.4.2 Dynamic Tests


The experimental setup for dynamic testing is shown in Figure 3.10. or in a photo in Figure 3.11.
It consisted of the dam monolith, USP and LSP and the springs to simulate hydrostatic loads.
The excitation of the model during the dynamic tests was due to shake table harmonic motions
of prescribed frequencies and amplitudes. A lateral restraining system consisting of three pairs
of rollers was used during all test in order to ensure motions of the dam model in the upstream
and downstream directions only.
The dynamic testing was conducted at frequencies from 5 to 20 Hz with a step of 2.5 Hz.
42

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

Figure 3.10: Experimental Setup for Dynamic Tests

Figure 3.11: Photo of Experimental Setup for Dynamic Tests

43

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

During each dynamic test, the shake table input motion time history was consisted of six parts,
all with the same frequency, but with successively increasing amplitude of acceleration. In the
first part, the amplitude was 75% of the maximum, i.e. 100% level, which was reached in the
sixth part. The amplitude difference between adjacent parts was steps 5% of the maximum
acceleration. The duration of shaking for each part was 10 seconds followed by 2 seconds of
rest to separate each part.
A typical dynamic test, for each combination of surfaces, hydrostatic force and frequency,
consisted of the following steps:

Set up the model on the shake table.

Attach horizontal long springs to downstream face of the model at specified height
and pre-stretch them sufficiently to provide a force equivalent to the hydrostatic
force due to the reservoir.

Conduct the test by exciting the model by prescribed motions of the shake table.
Measure the shake table acceleration, acceleration of the model at the bottom and
at the top, and the model displacement relative to the shake table. Maintain
hydrostatic force at constant level during the test. Shut down the excitations if the
model relative displacements are too large (about 15 cm) or i f the rate of the
model displacement is too high (about 10 cm in 10 sec).

Detach the horizontal springs.

Lift the model and clean the sliding surfaces.

At each frequency, the tests started at the peak acceleration levels of about O.lg. The levels

44

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

were increased in the subsequent tests until significant motions of the model were achieved.
After significant sliding was achieved an additional test for the same acceleration amplitudes
was repeated in order to verify the results.

3.5

INSTRUMENTATION

The instrumentation used for the static testing included two displacement transducers (Celesco
PT 101 Position Transducer), as shown in a photo in Figure 3.12, to measure relative
displacement of the dam with respect to the shake table. The compressive force between the
model and the rigid bar was measured by a load cell (Sensotec, model 41/0574-03) shown in a
photo in Figure 3.13. In addition, the force in the springs was also measured with a load cell of
the same type. The sampling rate for static tests was 20 samples per second.

Figure 3.12: Displacement Sensors Located on Downstream Side of the Monolith

45

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

Figure 3.13: View of a Load Cell to Measure Force between the Model and the Rigid Arm
The instrumentation used for the dynamic tests included two horizontal accelerometers (IC 3110
Accelerometers) to measure the upstream/downstream accelerations of the model close to the
base (see a photo in Figure 3.13), and at the top. The relative displacements of the model with
respect to the shake table were measured with two displacement sensors (Celesco PT 101
Position Transducer). Shake table horizontal motions were measured with an LVDT and an
accelerometer (Kistler 8304 K-Beam Accelerometer). The simulated hydrostatic force was
measured with a load cell (Sensotec, model 41/0574-03). The sampling rate for all dynamic
tests was 200 samples per second.

3.6 SUMMARY AND CONCLUSIONS


A number of static and dynamic tests were conducted during the preliminary experimental

46

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

study. These tests generated large amount of useful data, which included:

static and kinetic friction coefficients of the surfaces obtained from the static tests,

amplitudes of harmonic base accelerations to initiate sliding of the model at a


given frequency of excitations,
amplitudes of base accelerations to cause sliding of 3 cm during 10 seconds.

These were reduced, analysed and the results of this analysis were presented in a report
delivered to BC Hydro (Horyna et al., 1997). However, these results are not presented here
since the preliminary tests were superseded in many aspects by the further experiments, which
will be described in the next chapter with all results related directly to this thesis. Only a list of
general findings from the preliminary experiments is given here:

Combining several dynamic tests in one sequence reduced significantly testing


time of the dynamic tests and was found as a very convenient approach.

Response of the model during some tests contained significant undesirable out-of
plane rocking motions. These were caused by the fact that some surface plates
were not ideally flat and as a result of this the model to foundation contact was
limited to only several points. Resting on these points, the model rocked in crosscanyon direction during some tests. The restraining system, which was used
during the tests, only reduced the out-of plane rocking. This was considered when
the second phase of the tests was planned.

The hydrostatic load simulator used during the tests was able to provide desired
nearly constant horizontal force on the model. However, the springs had to be

47

CHAPTER

PRELIMINARY EXPERIMENTAL

STUDY OF A DAM MODEL

detached from the model before every moving it to the original position and after a
test and attached again before the next. Considering the number of tests needed to
conduct, this procedure was found to time demanding and other hydrostatic load
simulator was designed for the tests in the second phase.
A n attempt to simulate interlocking phenomena in the foundation interface was
done using the pair of plates with the Rough surface. However, this was found
very difficult and in order to reduce amount of uncertainties in the testing this was
removed as an objective of the second experimental phase.
A few experiments with records measured during past earthquakes were
conducted at the end of the testing. These tests generated useful information, but
it was found that scaling the records in time domain to obtain base excitations of
varied dominant frequencies caused complications during data analysis.
Therefore, instead of time-scaling records in the second phase it was proposed to
develop a series of synthetic records of the same duration but with different
dominant frequencies.
It was found that the sliding surfaces deteriorated partially during the series of
tests with one surface. In order to monitor the influence of this on the response of
the model in future tests, it was decided to perform static tests regularly after every
few dynamic tests of the next tests.

48

FURTHER EXPERIMENTAL
S T U D Y OF T H E D A M M O D E L

A series of shake table static and dynamic tests was described in the previous chapter. These
tests were conducted on the model of a concrete gravity dam monolith unbonded at the base. A
new series of shake table was scheduled then with similar goal as the preliminary tests with the
experience gained during those initial tests. As a result of this experience, several enhancements
of the experimental model, setup and testing procedures were introduced. Result of these
enhancements was that more types and a larger number of tests could be scheduled in the new
series.

These tests were conducted between May and July 1998 also in the Earthquake

Laboratory at U B C .

4.1 OBJECTIVES AND SCOPE OF FURTHER EXPERIMENTS


A total of four different types of tests were scheduled for this study. They included impact
hammer tests, static tests, dynamic tests with harmonic excitations (called harmonic tests) and
dynamic tests with synthetic earthquake excitations (called earthquake tests).

The objective of the harmonic and earthquake tests was to measure how much the experimental
model of a single monolith of a gravity dam slides if preloaded by a simulated hydrostatic force
and subjected to base excitations. The focus was put on frictional characteristics of the damfoundation interface and, therefore, the model was unbonded at the base during the tests. Other
forces in a real dam-foundation interface, such as cohesion, were not simulated.

Another objective of these tests was to study changes in the response of the model with varying
dominant frequency of applied base excitations. In particular, it was of interest to determine if
49

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

the range of testing frequencies can be sorted according to qualitative changes in the response
of the model.
Along with the dynamic tests, a series of static tests were scheduled. The objective of these tests
was to determine the static and kinetic coefficients of friction of the model-foundation interface
and static forces necessary to calculate the force ratios.
The objective of the impact hammer tests was to determine the natural frequencies of the
experimental model.
Similarly to the preliminary tests, the further experiments were not conducted in order to
simulate response of any specific prototype dam to an earthquake.

4.2 EXPERIMENTAL MODEL


The experimental model used during the preliminary tests was also employed during the
additional tests, except for two main differences:

Different shape and finishing of the upper surface plates, and

Different hydrostatic load simulator.

4.2.1 Sliding Surfaces


The sliding surfaces, fabricated by an external contractor, were designed to be in contact only
at the toe and heel zones of the dam monolith model. This was done in order to minimize outof plane rocking of the model during the shake table tests, which was observed during some of
the preliminary tests. The contact between the Upper Surface Plate (USP) and the Lower
Surface Plate (LSP) was limited to 150 mm at each end. This type of contact between the sliding

50

CHAPTER

FURTHER EXPERIMENTAL STUDY OF THE DAM

MODEL

surfaces was achieved by specifying raised sections at the end of each USP, which can be seen
in Figure 4.1. The shape of the LSP was the same as during the preliminary tests.
L

1165

00
<

Section B-B

1 0

Note: all dimensions are in mm, unless otherwise specified.


Figure 4.1: Schematic of the Upper Surface Plate

The Smooth surface was created by sanding the cement milk from the friction surface of the
plates to give a uniform surface. The Rough surface was developed using an ultra-high pressure
water jet to remove the cement matrix leaving an exposed aggregate surface. The water jet had
a pressure of 207 M P a and preparation of the Rough surface is shown in Figure 4.2. The
finishing was done on both contact surfaces, that is the top of LSP and the bottom of USP. A
51

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

total of four pairs of the surface plates were prepared, two with Smooth and Rough surface.

Figure 4.2: Photo Showing High Pressure Jet Preparing Rough Surface

4.2.2

Hydrostatic Load Simulation

The hydrostatic load was simulated by applying a pulling force in the downstream direction on
the downstream face of the dam model. This force was provided by a steel cable attached to the
model at a height corresponding to the resultant of the simulated hydrostatic load. The other end
of this cable was attached to a hanging weight equivalent to the required force. Figure 4.3 shows
a schematic of the hydrostatic load assembly and Figure 4.4 presents a photo including the
experimental model with the applied simulated hydrostatic force. The weight was attached to
two vertical rods which constrained it to move only in the vertical direction. Friction between
the weight and the sliding rods was minimized by using bronze-oilite bearings, which provided
a smooth contact surface.
52

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

During trial shake table harmonic tests with the assembly, the model was subjected to
undesirable significant force pulses introduced by the motion of the mass bouncing on the steel
cable attached to the model. These pulses manifested themselves when sliding of the model was
initiated. Since these prevented a constant level of simulated hydrostatic load, it was decided to
minimize their effects. This was accomplished by inserting a 25 cm long piece of rubber,
between the model and the cable.

This rubber piece, acting as a base-isolator, reduced

effectively the pulses on the model to negligible levels.


The simulated hydrostatic load was set to 1,040 N , which corresponded to a 95% of full
reservoir level considering the maximum water level as 97% of the total height of the model
including the USP. When this force was calculated, mass density of water was reduced by factor
S = 3.22, which corresponded to the ratio of the mass density of concrete to the mass density
p

of the material of the model and was calculated in the previous chapter. Although it was
desirable to keep this force constant during all tests, the resulting measured force fluctuated
somewhat during the tests as well as between tests. However, these variations did not exceed
5% during any of the tests.

4.3 INSTRUMENTATION
During the testing a maximum of ten different time history signals were recorded in each test.
The location of the instruments used to record the signals are shown in Figure 4.5. The arrows
labelled 6 through 10 show the location of triaxial accelerometers (IC 3110 Accelerometers),
which measured the horizontal and vertical accelerations of the model. The relative
displacement of the model with respect to the shake table (arrow 3 in Figure 4.5) was measured
54

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

Figure 4.5: Location of Instrumentation


using a displacement transducer (Celesco PT 101 Position Transducer).

The simulated

hydrostatic force (arrow 5 in Figure 4.5) was measured with a 11.1 k N load cell (Interface) and
the force applied by the arm, for the static-push tests, was measured with a 44.5 k N load cell
(Sensotec). Shake table horizontal displacement and acceleration (labelled 1, 2 respectively)
were measured with sensors built in the shake table (LVDT and a Kistler 8304 K-Beam
accelerometer). A complete list of the sensors used for the testing is given in Table 4.1. A l l
recorded signals were filtered with a 30 Hz low-pass filter using a 3 pole Bessel type filter with
a 60 dB/decade roll off.

55

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

Table 4.1: List of Instrumentation and Location


Location

Instrument

Displacement sensor

2
3
4
5

Accelerometer
Displacement
transducer
10 kip load cell
2.5 kip load cell

Accelerometer

Accelerometer

Accelerometer

Accelerometer

10

Accelerometer

Model

Measurement

Shake table horizontal


LVDT
displacement
Kistler 8304 K-Beam
Shake table horizontal
Accelerometer
acceleration
Celesco PT 101
Model displacement
Position Transducer
relative to shake table
Sensotec
Force in rigid arm
Interface
Simulated hydrostatic force
Model horizontal accel.,
u/s face, bottom
IC3110
Model vertical accel.,
IC3110
u/s face, bottom
Model horizontal accel.,
IC3110
d/s face, bottom
Model vertical accel.,
IC3110
d/s face, bottom
Model horizontal accel.,
IC3110
d/s face, top

4.4 TEST SETUP AND PROCEDURES


4.4.1 Impact Hammer Tests
The objective of the impact tests was to obtain the frequencies corresponding to the fundamental
modes of vibration of the experimental model. It was desirable to know if any of the frequencies
of excitation (harmonic and earthquake tests) was close to one of the natural frequencies of the
model. In addition, these tests were conducted in order to generate data for calibration of finite
element models developed during the study.
It was decided to perform the impact hammer tests for two types of boundary conditions:

unbonded (free-standing) model, and

bonded (fixed at four corners to the base) model.

56

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

Figure 4.7: View of Downstream Side of the Model Attached to Rigid Floor
during Impact Hammer Tests
57

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

Since the floor at the Earthquake Laboratory was not suitable for attaching the model to it, the
impact hammer tests were conducted at the U B C ' s Structures Laboratory. A total of 16 tests
were conducted in June, 1998. During half of these tests the model was fixed to the strong floor
of the lab at four corners (bonded model). View of the downstream side of the model fixed to
the floor is shown in Figure 4.7. The other half of the tests were done with the model freestanding on the floor (unbonded model) in order to monitor changes of its natural frequencies
for different boundary conditions. Only one Rough surface (Rl) was used for the testing.
A n instrumented sledge hammer (Dytran model 5803A) was used for the testing. The hammer
impacts were applied horizontally at the top and mid height of the model, on its upstream side.
Hard and soft tips of the hammer were used for the testing. The hydrostatic force was not
simulated during these tests. The data measured during the tests included signals from the
hammer and four horizontal accelerations measured with accelerometers (IC, model 3110) on
the upstream side of the model, at the top, bottom and thirds of the height of the model (see
Figure 4.6). Each record had a duration of 10 seconds and the measured signals were recorded
with a sampling rate of 500 samples per second.

4.4.2 Static Tests


The objectives of the static tests were to determine friction coefficients for each surface and to
track the deterioration of the surfaces as testing progressed. For the static tests the model was
held stationary while the table moved slowly under it. A n arm attached to a strong column
provided the restraint necessary to hold the model. Figure 4.8 shows a schematic of the arm and
load cell configuration. For the static tests the force in the strong arm, simulating the hydrostatic

58

CHAPTER

FURTHER

EXPERIMENTAL

STUDY

OF THE DAM

MODEL

Strong Column

2000

Figure 4.8: Schematic of Strong Arm Used For Static Tests

Figure 4.9: Photo of Typical Static Test Setup

59

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

force, shake table displacement and model displacement were measured (measurements 1, 3,4
and 5 in Figure 4.5) for 160 seconds at a sampling rate of 100 samples per second. Figure 4.9
shows a photo taken during one of the static test.
For each of the four surfaces (2 Smooth and 2 Rough), a number of preliminary static tests were
conducted. Relatively large changes in the force required to push the model were observed
during these first few tests. It was observed that it was during these initial tests that small
irregularities on the surface were removed or reduced as sliding between plates occurred.
Through the process of wearing down these irregularities it was also observed that the friction
coefficients remained stable and thus lifting and cleaning the contact surfaces between tests was
deemed unnecessary.
A series of preliminary static tests were conducted every time a new pair of plates was used for
the first time, and a single test was conducted between each change of excitation frequency
during harmonic and earthquake tests. Because the friction values were most stable when the
model was not lifted between tests, a test sequence was developed which would allow the model
to be pulled back up-stream between tests without lifting it up. This was accomplished by
means of a winch attached to the base of the model and the strong column, shown in Figure 4.10.
A typical static test consisted of the following steps:

Position the model on the shake table with simulated hydrostatic load applied.

Lower strong arm to horizontal position.

Extend strong arm to make contact with the dam model.

Conduct test (table slowly moves toward strong arm; model held stationary).
Pull model back to original position with the winch.

60

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

Figure 4.10: Winch Assembly Used to Pull Model Up-Stream Between Tests

4.4.3 Tests with Harmonic Input


The objective of the tests with harmonic input was to measure response of the dam monolith
model, preloaded by simulated hydrostatic force, due to harmonic base excitation. The
parameters varied during these tests included frequency and amplitude of the base acceleration.
Nine different frequencies were selected for the harmonic inputs. These frequencies were 5,7.5,
10, 12.5, 15, 17.5, 20, 22.5, and 25 Hz. As an example of an input signal, the 5 Hz harmonic
input signal is shown in Figure 4.11. This signal has 6 segments with different amplitudes, in
particular ranging from 0.525g to 0.7g, with increments of 0.035g. This record was used as the
input signal to the shake table where it was adjusted to the desired amplitude levels. At each
particular frequency the simulation was run several times, increasing the amplitude of shaking
each time, in order to capture the amplitude of motion which caused the model to slide a

61

CHAPTER

10

20

30

FURTHER EXPERIMENTAL

40
50
Time (sec)

STUDY OF THE DAM MODEL

60

70

80

90

Figure 4.11: Example of Harmonic Input Record at 5 Hz


prescribed amount. Typically, the record for a certain frequency was run 5 to 10 times providing
30 to 60 different ten second long records at different amplitudes.
During the harmonic tests, records from locations 1, 2, 3 and 5 through 10 in Figure 4.5 were
obtained. The duration of all tests was 90 seconds and each channel was recorded at a sampling
rate of 200 samples per second. A typical harmonic setup is shown in Figure 4.12.
A typical harmonic test included the following steps:

Position the model on the shake table with simulated hydrostatic load applied.

Run the test using a record with low amplitude (about 0.3g).

Increase amplitude of the record (by about 0.05g).

Continue increasing amplitude - pull back model to start position as needed.

Increase amplitude until model moves full distance in one testing segment.

Run Static Test.

4.4.4 Tests with Synthetic Earthquake Input


The objective of the earthquake tests was to measure response of the dam monolith model

62

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

Figure 4.12: Typical Harmonic Test Setup


preloaded by simulated hydrostatic force to a transient base excitation. The parameters varied
during these tests were the same as those for harmonic tests, the experimental setup and data
collection for the earthquake tests were also the same.
The records used for these tests were generated using the program S I M Q K E (Gasparini and
Vanmarcke, 1976) from the Power Spectral Densities (PSD) of recorded motions from three
selected past earthquakes in California. These records were:

1992 Landers earthquake, recorded at the Joshua Tree Fire Station, E/W direction
(CSMIP, 1992). This is designated as EQ1 for the remainder of this thesis;

1994 Northridge earthquake, recorded at the Tarzana Nursery Station, E/W


direction (CSMIP, 1994), designated as EQ2;

1979 Imperial Valley earthquake, recorded at El Centro, Bonds Corner (Highways


98 and 115) Station, SW/NS direction (California Department of Conservation,
63

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

1983), designated as EQ3.


The records used for these tests were developed as follows. First, the PSD from each of the
above records was calculated and smoothened using a 50 point moving average. Then, the
resulting PSD values were shifted so that the frequency of the maximum PSD value
corresponded to the frequency of interest. This frequency was considered to be the dominant
frequency fj of the generated record. The shifted PSD values were multiplied by a window
function of frequency band (7^-2) Hz to (/^-4) Hz in order to narrow-band the record being
developed into a 6 Hz frequency band. The 6 Hz band was selected after it was observed that
all the original PSD's had negligible values out of this band. Finally, program SIMQKE was
used to generate a synthetic record with a duration of 10 seconds. This process was repeated
over 9 dominant frequencies (from 5 to 25 Hz with a step of 2.5 Hz) using each of the 3 seed
earthquake records (EQ1, EQ2 and EQ3). Each of the generated records were identified by
reference to the name of the seed earthquake and by the value of the dominant frequency. For
example, synthetic earthquake EQ1 with a dominant frequency of 12.5 Hz had components with
frequencies from 10.5 H z to 16.5 Hz, and it was derived from the Landers earthquake seed
record. Generated records from the three different seed records and with the same dominant
frequency were combined into one time history, which was later used as a driving signal for the
shake table. A n example of such driving signal is shown in Figure 4.13, which presents three
synthetic earthquake records, each with a dominant frequency of 5 Hz.
A typical earthquake test was as follows:

Position the model on the shake table with simulated hydrostatic load applied.

Run the simulated earthquake record with low amplitude (around 0.3g).

64

CHAPTER

FURTHER EXPERIMENTAL

10

20

STUDY OF THE DAM MODEL

30

40

Time (sec)

Figure 4.13: Example of Synthetic Earthquake Input Record with Dominant Frequency 5 Hz

Increase amplitude of record (add 0.05g).

Continue increasing amplitude - pull back model to start position as needed.

Increase amplitude until model moves full distance in one test segment.

Run Static Test.

4.5 RESULTS OF EXPERIMENTS


A total of 360 harmonic tests, about the same amount as the earthquake tests, and more than 100
static tests were conducted using the shake table. Nearly 1 GB of data was generated by the tests
and it was necessary to reduce, analyse and evaluate all these data.
This section presents the main findings obtained from the experimental work. Out of the tests
conducted on all four surfaces (Rough surfaces R l and R2, and Smooth Surfaces SI and S2),
the experimental results presented here are limited to those from the tests conducted on one
surface, the Rough surface R l .

This is because all numerical analyses of the experimental

65

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

model described in the next chapter of this thesis were performed using the frictional
characteristics of this surface. The test results from other surfaces are not presented here. A
complete set of results from all surfaces can be found in Black et al. (1998).

4.5.1 Results of Impact Hammer Tests


The impact hammer testing was conducted with the bonded model, that is attached to the strong
floor at the U B C ' s Structures Laboratory and with the unbonded model, that is free-standing on
the floor. The results of the tests in terms of the natural frequencies and characteristics of the
corresponding mode shapes are presented in Tables 4.2 and 4.3. The natural frequencies were
determined from peaks of the Frequency Response Function (FRF) considering the signals from
the hammer as the input and those from the accelerometers as the output of a single degree of
freedom system (Bendat and Piersol, 1971). From a comparison of results shown in Tables 4.2
and 4.3, it can be observed that the natural frequencies of the bonded model are quite different
from those of the unbonded, free-standing, model. For example, the fundamental natural
frequency of the bonded model is about twice as high as that of the free-standing model.
It also can be observed from the values in Tables 4.2 and 4.3 that the variability in the frequency
values are different for both models. The bonded model exhibited approximately the same
natural frequencies for a wide range of tests. In contrast, the unbonded model had natural
frequencies dependent on various parameters such as the power of the hammer strike, location
of the strike, and the position of the model on the base. These variations in natural frequencies
of the unbonded model can be attributed to nonlinearities of the setup, such as rocking of the
model and slight changes of the contact characteristics at the foundation interface.

66

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

Table 4.2: Natural Frequencies of the Bonded Model


Frequency (Hz)

Characteristics of Natural Mode

66

the 1st cantilever bending mode


(no node along the model height)

174

the 2nd cantilever bending mode


(one node along the model height)

Table 4.3: Natural Frequencies of the Unbonded Model


Frequency (Hz)

Characteristics of Natural Mode

27-34

rocking combined with the 1st cantilever bending mode

53-63

character of the mode could not be identified from the


tests (such identification is carried out in Chapter 5
with the help of FE model)

140

second cantilever mode (one node along height)


combined with rocking

The coordinates of the natural modes were determined using amplitudes and phase angles of the
FRF calculated from the measured signals. However, because the model was instrumented with
four horizontal accelerometers only, it was not possible to identify precisely the characteristics
of the natural modes of the free-standing model as these modes exhibited components in the
horizontal and vertical directions. Results of a finite element analysis of the model conducted
simultaneously by Rudolf (1998) were used as complement to the experimental results to
identify the character of the natural modes of the free-standing model.
The impact tests were also used to obtain damping of the system. This was determined from the
free vibration decay of amplitudes of horizontal acceleration measured at the top of the
monolith. The damping was found to be about 5% of critical for the first mode of vibration.
67

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

4.5.2 Results of Static Tests


A number of static tests were conducted on each surface prior to any harmonic or earthquake
tests in order to condition the surface before each series of tests. Results from these trial tests
were not used for any analysis. Static tests were conducted before every series of the harmonic
or earthquake tests with a single dominant frequency, in order to determine the friction
coefficients prior to that series.
A n example of measured time histories during the static tests is shown in Figure 4.14. The
quantities plotted in this figure are identified above each figure. It can be observed from the plot
of the total force on the model (part c in the figure) that the friction characteristics of the surface
RI were almost uniform along the entire tested path about 8 cm long (see part d). It can be also
concluded that no significant drop of the friction force was observed after the sliding started,
which means that the difference between the static and kinetic friction was small for surface R I .
The results of static tests in terms static and kinetic friction coefficients are given in Table 4.4.
The friction coefficients were determined from the signal obtained by adding the measured force
time history, obtained from the load cell between the strong arm and strong column, and the
measured simulated hydrostatic load. The static coefficient was calculated as the average of the
peaks in this signal divided by the weight of the model, while the kinetic coefficient was
obtained as the average trough of the signal divided by the weight.

68

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

Figure 4.14: Example of Signals Recorded during Static Tests with Surface R l
69

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

Table 4.4: Friction Coefficients of Rough Surface R l Obtained from Static Tests
Test#

Measurement Taken Prior to:

Static Friction
Coefficient

Kinetic Friction
Coefficient

16

Harmonic test at 5 H z

0.73

0.72

17

Harmonic test at 7.5 Hz

0.72

0.71

18

Harmonic test at 10 Hz

0.74

0.73

19

Harmonic test at 12.5 H z

0.74

0.73

20

Harmonic test at 15 Hz

0.74

0.74

21

Harmonic test at 17.5 Hz

0.75

0.75

22

Harmonic test at 20 Hz

0.75

0.74

23

Harmonic test at 22.5 Hz

0.76

0.76

24

Harmonic test at 25 H z

0.76

0.75

28

Earthquake test at 5 Hz

0.75

0.75

29

Earthquake test at 7.5 Hz

0.76

0.75

30

Earthquake test at 10 Hz

0.76

0.75

31

Earthquake test at 12.5 Hz

0.76

0.75

32

Earthquake test at 15 Hz

0.77

0.76

33

Earthquake test at 17.5 Hz

0.76

0.75

34

Earthquake test at 20 Hz

0.77

0.76

35

Earthquake test at 22.5 H z

0.77

0.77

36

Earthquake test at 25 Hz

0.77

0.77

The differences between the static and kinetic friction coefficients obtained from the same test
are very small. No significant fluctuations in the friction properties of this surface were found
between the tests. The average static friction and kinetic friction coefficients were determined
as mean values of the third and fourth columns of Table 4.4 as 0.75 and 0.74, respectively.
However, it can be observed that there was an obvious trend of the friction coefficients
increasing as the testing progressed. The increase in the friction coefficients as a result of

70

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

surface changes during the trial tests had been taken into account in further analysis of the
experimental data and in the numerical study to simulate some of the shake table tests.

4.5.3 Results of Tests with Harmonic Input


The data measured during the harmonic tests was converted into a form suitable for analysis and
some basic signal processing was performed. An example of signals after this analysis is given
in Figure 4.15. showing shake table motions and selected response of the experimental model
during one of the harmonic tests. Each of plotted quantities is identified above its graph. The
plots cover a sequence of six test segments as these were conducted during the test. It can be
observed from the plots how the sliding of the model increased with increasing amplitudes of
the shake table motions. The plot of the simulated hydrostatic force indicates that it exhibited
some minor fluctuations when the sliding of the model occurred, but these fluctuations were
found small and they did not exceed 5% during any of the tests.
Reduced results of the data obtained from the harmonic tests are given in Figure 4.16. This
contains 9 plots of the measured Peak Table Accelerations (PTA) and Root Mean Square (RMS)
acceleration versus Rate of Model Displacement (RMD), one plot for each testing frequency.
The PTA was obtained as a maximum absolute value of a base acceleration record from every
test. The R M S acceleration a

was calculated as a Root Mean Square of a vector containing

RMS

positive and negative peaks of the base acceleration record:


'
RMS

JjZ\P
N

j\

eak

(4.1)

7=1

where N is the number of the peaks greater than 50% of the maximum absolute acceleration in
71

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

a) Shake Table Displacement

-1
2

b) Shake Table Acceleration

1
0
-1
-2
2

c) Model Base Acceleration

1
0
-1
-2

1.5
2

.
1
1 1 1 1 ,
d) Simulated Hydrostatic Load

1.0
0.5
I

0.0

e) Displacement of Model Base

10

20

30

40

50

Time (sec)

60

70

80

90

Figure 4.15: Example of Signals Recorded during Harmonic Tests with Surface R l
72

CHAPTER

FURTHER EXPERIMENTAL

b) 7.5 Hz

a) 5 Hz
2.0

"ir

1.5

c) 10 Hz

2.0

2.0

1.5

1.5

o 1.0

o 1.0

8o
<

<

<

0.5

0.5

0.0

0.0

0.0

10

15

20

RMD (mm/s)

10

15

20

f) 17.5 Hz

2.0

o> 1.5

1.5

~ro1 5 h

1.0

a5 1-0

0.5

<

o 1.0
0.5
0.0

t
0

4 *

Q)
O
O
<

20

2.0
_
3

'

1.5

8
o
<

1.0

r~

0.5
0.0

20

5
10 15
RMD (mm/s)

2.0

1.5

1.5

1.0

20

i) 25 Hz

2.0

o
<

0.5

0.5
0.0 0

15

h) 22.5 Hz

r~

<

10

RMD (mm/s)

g) 20 Hz

2.0

oo

0.0
5
10 15
RMD (mm/s)

5
10 15 20
RMD (mm/s)

e) 15 Hz

2.0

<

RMD (mm/s)

d) 12.5 Hz

1.0

0.5

STUDY OF THE DAM MODEL

io

0.5
5
10
RMD (mm/s)

0.0
-10

-5
0
5
RMD (mm/s)

10

0.0-10

-5

10

RMD (mm/s)

RMS
PTA

Figure 4.16: Table Acceleration vs. Rate of Model Displacement for Harmonic Excitation
73

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

the signal and peakj is the j-th element of this vector. When calculating the

value, the

peaks lower than 50% of PTA were disregarded because it was assumed that these do not
contribute to the sliding of the dam model.
The concept of R M D was introduced from the following reason: During some tests, when the
model was approaching the limit available for its sliding, the base excitations had to be suddenly
terminated in order to protect certain parts of the model from permanent damage. As a result of
this, there were several tests containing valuable information, but the tests were of different
duration than the nominal. R M D was introduced to eliminate the duration of the test in further
analysis of the data and it was calculated as a ratio of the measured displacement divided by the
actual duration of the test.
The R M S acceleration values are included together with the P T A values in Figure 4.16.
Associating an R M D to a single measured P T A , which in some cases may be
uncharacteristically high, may not be the best way to characterise the record. The RMS value
may provide better information about the average levels of acceleration that resulted in sliding
of the model. Further reduction and analysis of the data from harmonic tests is provided later
in this section where these data are compared with that obtained from the earthquake tests.

4.5.4 Results of Tests w i t h Synthetic Earthquake Input


The tests with the earthquake records yielded significant amount of data, which were analysed
and interpreted in several stages. This section describes basic reduction of these data.
The records measured during the tests were converted to a format suitable for analysis,
conditioned and visually checked for possible errors. A n example of time histories describing
74

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

shake table motions simulating three synthetic earthquakes and response of the model due to
these motions is shown in Figure 4.17. The plotted quantities are identified above each graph.
Analysis of the data obtained during earthquake tests, to determine relationships between shake
table accelerations and R M D values, was performed in a way similar to that for the harmonic
tests. It was considered that simply taking the measured PTA could be misleading as it might
be a single acceleration spike. Obviously this spike should not be the parameter or the effective
table acceleration used to characterize the acceleration, which caused certain amount of
displacement. To alleviate this problem, two versions of the table acceleration were used: the
actual measured peak table acceleration (PTA); and 2) the R M S of the peaks in the acceleration
time history. When calculating the R M S value, only those peaks greater than 50% of the PTA
were considered, while the peaks lower than 50% of PTA were disregarded as in the case for
harmonic excitations. Every test segment from the number of earthquake tests conducted was
analysed to obtain the measured R M D corresponding to a given table acceleration (PTA or
RMS). These results are given in Figures 4.18 to 4.20, which show the acceleration versus
R M D for each earthquake and surface combinations.
Similarly to that of the harmonic tests, further reduction and analysis of the data from earthquake
tests is not performed independently. Instead, the combined information from Figures 4.17 to
4.20 is analysed in the next subsection in order to get direct comparison from tests with different
base excitations.

75

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

a) Shake Table

10

20
Time (sec)

30

40

Figure 4.17: Example of Signals Recorded during Earthquake Tests with Surface RI

76

CHAPTER

FURTHER EXPERIMENTAL

a) 5 Hz
i

++

b) 7.5 Hz
2.0

STUDY OF THE DAM MODEL

_1.5
3

8 LO

c)10Hz
~i i

2.0

_1.5
^

++

<

LO

* " "

++

0.5
J

10

L_

15

20

0.0 0

RMD (mm/s)

0.0

15

1.5

2.0

1.5

1.5

5
10 15
RMD (mm/s)

0-5

0.0
20

1.5

' ' '

10

15

0.0
20

5
10 15
RMD (mm/s)

h) 22.5 Hz

1.0

RMD (mm/s)

g) 20 Hz
2.0

0.5

2.0

1.5

1.5

1 0

1.0

0.5

0.5

......

-k
+
A

i
0

20

i) 25 Hz

2.0

0.0

20

3
"55 1.0

0.5

r
0

15

++

10

l_

f) 17.5 Hz

2.0

1.0

0.0

RMD (mm/s)

t
++

e) 15 Hz

% 1.0

20

RMD (mm/s)

d) 12.5 Hz
2.0

10

10

0.0
15

RMD (mm/s)

20

-10

-5
0
5
RMD (mm/s)

10

-10

-5

10

RMD (mm/s)

RMS

A PTA

Figure 4.18: Table Acceleration vs. Rate of Model Displacement for Excitation EQ1
77

CHAPTER

FURTHER EXPERIMENTAL STUDY OF THE DAM MODEL

a) 5 Hz
2.0

-1.5

_1.5
3

"i r

1.5

O)

1.0
<

A*

0.5

0.0

0.0
0

10

15

20

<
0

10

15

" 1.0
<
8

o
< 0.5

A - "

f) 17.5 Hz

20

_
S

10

15

0.0-10

0.0-10

-5
0
5
RMD (mm/s)
+
A

' A ......
A

*t,

20

i) 25 Hz

1.5

0.5 (
10

5
10 15
RMD (mm/s)

2.0

+.A

4*--

-5
0
5
RMD (mm/s)

...+..;..*..

1.5

<

20

AM.

3
1.0
o

h) 22.5 Hz
2.0

1.5

0.0

RMD (mm/s)

2.0

0.5

<

o5 1.0
o
o

g) 20 Hz

<

OT

0.0

5
10 15
RMD (mm/s)

5
10 15 20
RMD (mm/s)

2.0

i
0

... A.:

1.0

0.0

+
, 1

1.5

Q)

tr

'4T

0.0

1.5

+ :

e) 15 Hz
2.0

ra 1.5

0.5

*-++

20

RMD (mm/s)

d) 12.5 Hz

1 1> i

A
A

0.5

RMD (mm/s)

2.0

0.5

t 1-0

<

c) 10 Hz
2.0

+.

Acce

b) 7.5 Hz
2.0

HA

0.5

4
.. . J . .i .
0.0

10

-10

-5

10

RMD (mm/s)

RMS
PTA

Figure 4.19: Table Acceleration vs. Rate of Model Displacement for Excitation EQ2

78

CHAPTER

FURTHER EXPERIMENTAL STUDY OF THE DAM MODEL

a) 5 Hz

b) 7.5 Hz

2.0

2.0

1.5

O)

'

c) 10 Hz
2.0

1.5

'

1-5
D)

A A A

1.0

'

1.0

1.0

p
0.5

0.5

0.5
--

0.0

0.0
0

10

15

20

RMD (mm/s)

0.0

10

15

20

5
10 15 20
RMD (mm/s)

RMD (mm/s)

f) 17.5 Hz

e)15Hz
1

'

2.0

~1

1.5

3
^1.0

0.5

j
i . i .

5
10 15
RMD (mm/s)

20

-ii

0.0

10

15

20

5
10 15
RMD (mm/s)

RMD (mm/s)

g) 20 Hz
2.0

i4

h) 22.5 Hz
2.0

i) 25 Hz

Jk

1.5

1.5

+:
4-

O)

<

1.5

1.0

1 0

20

1.0

A +

0.5

0.0 0

0.5
>

10

RMD (mm/s)

0.0

0.5
. . . i . i
-10

0.0

-5
0
5
RMD (mm/s)

+
A

10

-10

-5

10

RMD (mm/s)

RMS
PTA

Figure 4.20: Table Acceleration vs. Rate of Model Displacement for Excitation EQ3
79

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

4.5.5 Comparison of Results from Harmonic and Earthquake Tests


In the previous two subsections the data from the harmonic and earthquake tests was reduced
into the form of R M D to PTA/RMS relationships and these were presented in Figures 4.16 and
4.18 to 4.20. Based in this information and notes and video tapes taken during the tests, the
testing frequencies can be divided into three groups depending on the response of the model.
These groups can be characterized as follows:

Group 1 - low frequencies: The model slid when the friction was overcome; no
in-plane rocking, that is the rocking about an axis in cross canyon direction, was
not observed and; resulting sliding of the model after the tests was always
downstream - see Figures 4.16 a and b and 4.18 to 4.20 a and b.

Group 2 - medium frequencies: Sliding of the model was affected by flexural


behaviour and rocking of the model. As a result of this the sliding of the model
was initiated at lower amplitudes than it was for the frequencies from Group 1.
The amounts of sliding measured for the frequencies from Group 2 were mostly
larger than those for the same amplitude of base acceleration and frequency of
excitation from Group 1 - compare Figure 4.16 e and f with a and b, or Figure
4.20 e and f with a and b.

Group 3 - high frequencies: The primary response of the model to the base
excitation, which could be clearly observed and heard, was rocking. The resulting
sliding of the model depended on both the frequency and amplitude of base
excitation. Downstream, upstream as well as almost no motions were observed,
however the upstream motions dominated especially for high amplitudes of base
accelerations. In some cases, the model started to slide downstream, but at higher
amplitudes of base acceleration it slid upstream - see Figure 4.18 or 4.19 h and i.
Using the video tapes taken during the tests and from the visual observation of the

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

model's behaviour during the tests, it was concluded that the model, during the
tests at frequencies from 20 to 25 Hz combined with certain amplitudes of base
accelerations, was not in contact for short periods of time. If the model underwent
such motions while the table moved downstream, the model moved slightly
upstream relatively to the table. When such relative motions were cumulated they
resulted in the residual upstream sliding of the model.
Using the above grouping scheme applied on the results from harmonic and earthquake tests on
surface RI the information given in Table 4.5 was developed. It can be concluded from this
table that the bounds of Groups 1,2, and 3 for each of the four testing surfaces are very similar
for all types of excitation. It seems that the effects of nonlinear phenomena, such as rocking,
manifested themselves at slightly lower frequencies in case of harmonic excitations than in case
of synthetic earthquakes. This can be explained because a single frequency excitation should
generate any kind of resonance sooner than more transient synthetic earthquake motion.
Table 4.5: Grouping Testing Frequencies According to Response of the Model
Excitation

Group 1

Group 2

Group 3

Harmonic

5 to 10 Hz

12.5 to 17.5 Hz

20 to 25 Hz

EQ1

5 to 12.5 Hz

15 to 20 Hz

22.5 to 25 Hz

EQ2

5 to 10 Hz

12.5 to 20 Hz

22.5 to 25 Hz

EQ3

5 to 12.5 Hz

17.5 to 20 Hz

22.5 to 25 Hz

It can also be observed from Table 4.5 that the groups are, except for a couple of minor
irregularities, the same for all four types of excitation considered. Similar conclusions were
made from the results of tests using the other surfaces, which are not shown here. This means
that for different surfaces, the response of the model was controlled by the same phenomena and

81

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

it also means that the results of the tests were repeatable.


It can be concluded from the comparison of plots in Figure 4.16 with the corresponding ones in
Figures 4.18 to 4.20 that R M S to R M D relationships show a better agreement over the range of
base excitations used than the PTA to R M D relationships. The R M S and PTA values are about
the same for the harmonic excitations at each of the testing frequencies, but they vary
significantly for all synthetic earthquake inputs. It appears that R M S is a better measure to use
than P T A in characterising a base excitation regarding sliding response of a structure. The
amount of sliding of such a structure does not depend only on intensity of a single acceleration
pulse, but it is a function of the intensity and the number of all acceleration pulses of the
considered earthquake.
It was shown by the experiments that the model slid upstream at frequencies of base excitation
from about 20 to 25 Hz. This is an important finding, but its practical significance should not
be overestimated. The frequency range of 20 to 25 Hz on a model frequency scale represents a
range of 5.2 to 6.5 Hz on a prototype 15-times larger and it is possible that the base excitation
acting on a prototype during an earthquake would have significant components at and above
these frequencies. However, it should be remembered that the experimental model used for this
study did not simulate several capacity (e.g. cohesion) and demand (e.g. hydrodynamic pressure
and uplift) components of a gravity dam-water-foundation system during an earthquake. To
reach a conclusion from this upstream motion finding, without confirmation using a model
including simulation of all the important phenomena, could be misleading. Such testing is
beyond the scope of this study.

82

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

4.6 SUMMARY OF EXPERIMENTAL WORK


Two series of experiments were described in Chapters 3 and 4.

These tests had several

objectives, which were listed at the beginning of the chapters.

In order to satisfy these

objectives, an experimental model of a single monolith of a concrete gravity dam was designed
and developed. Only a limited number of phenomena involved in the behaviour of a real damwater-foundation system were modelled. In addition, it was understood that validity of all the
tests is limited to the experimental model only and that it was not the objective of the tests to
predict performance of any real concrete gravity dam.
The model was subjected to a series of preliminary tests, during which important information
leading to several enhancements of the model, experimental setup and testing procedures was
collected. The second series consisted of the impact hammer, shake table static and shake table
dynamic tests. The shake table dynamic tests included those with the harmonic and synthetic
earthquake input.

Analysis of the data obtained from this series, directly related to the

objectives of this thesis, was performed in Chapter 4. This included reduction of data from all
tests and in addition to this:

Natural frequencies of the experimental model were extracted from data from the
impact hammer tests. These frequencies were summarised in Tables 4.2 for the
bonded model and in Table 4.3 for the unbonded model. Different values in these
tables indicate large influence of boundary conditions on the natural frequencies.
Static and kinetic friction coefficients of the tested model-foundation interface
were determined from the data obtained during the static shake table tests. These
were summarised in Table 4.4.
83

CHAPTER

FURTHER EXPERIMENTAL

STUDY OF THE DAM MODEL

Amounts of sliding of the experimental model due to harmonic or synthetic


earthquake base excitations were determined from the data gathered during shake
table dynamic tests. These results were presented in Figures 4.16 and 4.18 to 4.20.

Results of the shake table harmonic tests were compared with those using
synthetic earthquakes. It was found that the model responded in a very similar way
during both types of tests as can be seen from grouping the testing frequencies
according to the model's behaviour during the tests. This is shown in Table 4.5.

It should be mentioned that the amount of reported experimental data and the depth of the data
analysis presented here is limited just to ensure the continuity of the studies described in this
thesis. A more detailed analysis of the data gathered during the experimental study was
presented by Black et al. (1998). Some of the data obtained during the experiments will be used
in the next section for verification and calibration of numerical models.

84

A N A L Y T I C A L STUDY
OF THE DAM MONOLITH MODEL

A series of experiments on a scaled model of a gravity dam monolith was described in the
previous chapter. The development of three numerical models to simulate the tests is described
in this chapter. A large amount of data collected during the tests is used in this chapter to verify
and calibrate numerical models to simulate the shake table tests.
The main objective of the analytical study was to develop a simple numerical model to simulate
sliding of a rigid block on a rigid foundation, preloaded by a constant horizontal force and
subjected to base excitations. The requirement of the model's simplicity came from one of the
intended applications of the model. This was its use in reliability analysis, which typically
involves a large number of simulations and therefore, a fast analysis procedure is preferable.
Another objective of this chapter was to develop a numerical model more complex than the rigid
block model, and to use it to simulate some of the shake table tests conducted during the
experimental part of the study. Purpose of this analysis was to find out how closely could the
numerical models of varied complexity simulate the response of the experimental model
measured during the shake table tests.

5.1 DEVELOPMENT OF NUMERICAL MODELS


A total of three numerical models were developed during this study. They varied in their
complexity and consequently, in their ability to simulate desired phenomena accurately. In
addition, the models varied in computational effort necessary for a single simulation.
The following numerical models were developed:
85

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

Single Degree Of Freedom model - SDOF model;

3-Degree of Freedom model using software

Working Model (Knowledge

Revolution, 1996) - 3DOF model; and

model using finite element software A N S Y S (SAS IP, 1996) - F E model.

5.1.1 SDOF Numerical Model


The SDOF model, shown in Figure 5.1, consisted of a block with mass m resting on a rigid
foundation. The block was preloaded by a horizontal force F and subjected to a base motions
described by a function z(t). In the analysis of the system, it was convenient to describe motions
of the block in terms of its absolute displacement function y(t). The solution to the problem was
based on the following assumptions:

The block and the foundation are rigid.

Block is constrained to a single horizontal degree of freedom.

Frictional contact between the block and the foundation is assumed, with static
and kinetic friction coefficients. The static friction coefficient is applied i f the
relative velocity of the block with respect to the foundation is smaller than a
specified small number and the kinetic friction coefficient is applied i f the relative
velocity is larger or equal to that number.

The block and the base are always in contact, which means that any jumping or
rocking motions of the model is not considered.

The horizontal force F is constant and it is smaller than the friction force, which
could be transferred by the block-base interface. This force can be expressed as
86

CHAPTER

ANALYTICAL

STUDY OF THE DAM MONOLITH

MODEL

mg\i, where m is the mass of the block, p is the friction coefficient, and g is
gravity acceleration due to gravity.
No cohesion between the block and the foundation is considered.

Reference Plane

y(t)
Block

Base

Figure 5.1: Schematic of the SDOF Model

A closed form solution to a simplified problem was presented by Westermo and Udwadia
(1983). They showed solution to a simplified problem without the horizontal force F and
considering base motions limited to harmonic in nature. In this study, the solution of Westermo
and Udwadia was extended for the case when the horizontal force F is acting on the block and
the base excitation is of a general character. This is presented below.
It is assumed that the block does not slide at the beginning of the simulation (stick mode). In
such a case, the absolute acceleration of the block y(t) is the same as the base acceleration:
Rt)

= ho

In the stick mode, the velocities y(t)

(5.i)
and z(t) are equal, which means that the relative

displacement of the block with respect to the base does not change. The block remains in stick
mode until the time when the resultant of the inertia force my(t) and the force F exceeds the
87

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

friction force mg\i. In such a time, sliding of the block is initiated (slip mode) and its motion
is controlled by the equation obtained from the equilibrium of the horizontal forces on the block:
\y(t)m\ = mg\i-F

(5.2)

Depending on what is the sign of y(t) when the sliding was initiated (at time t ):
0

y\t) = p g - i f y\t )>0,and

(5.3)

y i t )

= -pg-if

;Kf )<0.

(5.4)

After the sliding is initiated, the absolute acceleration of the block at any time is given by one
of the Eqs (5.3) or (5.4). According to these, y(t) is a constant, which after integrating with
respect to time, leads to a linear function for absolute velocity y(t) and a quadratic function for
the absolute displacement y(t) of the block. The base velocity z(t) and displacement z(t)
have known values at any time. Thus, the relative velocity y {t) and the relative displacement
r

(sliding) y (t) can be calculated at any time:


r

yM

= An-kO

,and

(5.5)

y (t) = y(t)-z(0

(5.6)

The sliding will terminate, when the relative velocity of the block y (t) is equal to zero. At
r

such a time, the block will go to the stick mode. In this mode, the absolute motions of the block
will equal to the motions of the base. For base excitations with cyclic character, the above
described cycle typically repeats several time as the block is subjected to the base motions.
The solution described above implemented into a computer program, which was tested for the
block with the following parameters. The mass of the block was 480.6 kg, which corresponded

88

CHAPTER

ANALYTICAL

STUDY

OF THE DAM

MONOLITH

MODEL

to the mass of the experimental dam model used in this study. The static friction coefficient was
0.77 and the kinetic friction coefficient 0.74, which corresponded to the coefficients of the
Smooth surface from the preliminary experimental study, as can be seen in Table 7 in Horyna et
al. (1997). The magnitude of the force Fwas 1.7 kN, which corresponded to one of the values
of this force used during the preliminary experiments. The block was subjected to a 5 Hz
harmonic excitation, 1 second long. Four acceleration amplitudes were used for this study:
OAg, 0.6g, l.lg, and 1.4g. The solutions obtained using the SDOF model for these four cases
are shown in Figures 5.2 to 5.5.

Each figure contains four plots: acceleration, velocity,

displacement, and sliding. The base motions are plotted using dotted lines, the model absolute
motions using solid lines and the model relative motions with dashed lines.
The accelerations in Figure 5.2 indicate that sliding of the block was not initiated for OAg
harmonic excitations at 5 Hz. The response of the block to the 0.6g harmonic excitations is
shown in Figure 5.3. In this case, sliding of the block was initiated only during the negative
pulses of base acceleration, when the sense of the inertia and static forces on the model were the
same. The response of the block to excitations with amplitudes of 1 Ag, see Figure 5.4, shows
that sliding of the model was initiated in both cases, during the positive and the negative parts
of the cycle of base acceleration. Figure 5.5 presents the results of SDOF analysis for the
excitation with amplitudes of 1 Ag. In this case the sliding of the model was initiated during the
first pulse and it did not stop until the end of the excitation. This was due to the fact that the
relative velocity of the block with respect to the base never reached zero at the same time when
the demand on the system given by combination of the inertia and simulated hydrostatic forces
was smaller than the frictional capacity of the system.

89

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

0.001

fr 0.000
c

Block

-0.001

4E-3
Base
Block

2E-1

Block (relative)

5E-1

Block
Base

1.0
time (s)

Figure 5.2: Performance of SDOF Model - Harmonic Excitations at 5 Hz and 0.4g

90

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

0.010

cn
c
t75

Block

0.000 A

2E-2

Base

0E-+O

Block

8
ro
o.
w

i5

-2E-2

2E-1

Block (relative)

1E40

Block
Base

0.0

0.5

1.0
time (s)

1.5

2.0

Figure 5.3: Performance of SDOF Model - Harmonic Excitations at 5 Hz and 0.6g

91

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

0.200

c
jo

Block

0.000

2r>i

Base
Block

4E-1
Base
Block
Block (relative)

2E40

3
I

Block

OE40

Base

0
-2E-K)

0.0

0.5

1.0
time (s)

1.5

2.0

Figure 5.4: Performance of SDOF Model - Harmonic Excitations at 5 Hz and l . l g

92

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

0.400-,

Block

T3

0.000

4E-1

OE40

Base
Block

-4E-1

Figure 5.5: Performance of SDOF Model - Harmonic Excitations at 5 Hz and 1.4g

93

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

It can be concluded that the SDOF model can simulate sliding of a rigid block preloaded by a
constant horizontal force due to base excitations. The level of accuracy of the results using this
model will be discussed in the next section.

5.1.2 3DOF Numerical Model


Working Model (WM) is a tool for engineering simulation. It was developed by Knowledge
Revolution of San Mateo, California, U S A . For this study, a 2D version 4.0 was used
(Knowledge Revolution, 1996). The program's graphical user interface allows the user to
define a set of rigid bodies and constraints and simulate its behaviour using a Newtonian
mechanics simulation engine.
The 3DOF numerical model included the shake table, the dam monolith model and an actuator
to simulate the hydrostatic load on the model. The shake table was constrained to move in the
horizontal direction only and it was powered by an acceleration controlled actuator. The
horizontal actuator simulating the hydrostatic force on the model, was force controlled. It was
stretched between the downstream side of the dam model and a support fixed to the shake table.
The model of the dam monolith had three degrees of freedom, allowing the centre of gravity of
the model to move horizontally and vertically as well as rotate. This means that the 3DOF
model could account for sliding and rocking of the model and the frictional contact between the
model and its base may not exist at certain times of the numerical simulation. The solution to
the system of three equations of motion, corresponding to the three degrees of freedom, was
obtained using the W M ' s simulation engine, using a Kutta-Merson (5th-order Runge-Kutta)
time integration scheme in order to increase accuracy of the solution.

94

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

a) Harmonic Excitation at Frequency 5 Hz


1.0

n
i

0.8
0.6

<
OH

I
i

+ SDOF

0.4
;

0.2
0.0

3DOF

i
5

10

i
15

20

25

30

RMD (mm/sec)
b) Harmonic Excitation at Frequency 20 Hz
1.0

~i
i

0.8

0.6

<
CL,

0.4
0.2

. 3DOF
+ SDOF

|
!

0.0
0

10

15

20

25

30

RMD (mm/s)
Figure 5.6: Comparison of Results from 3DOF and SDOF Models for Harmonic Excitations
The 3DOF model was developed before the results from the further experiments were available
and therefore it was tested against model's parameters corresponding to those used in the
preliminary tests (Horyna et. al, 1997), which were also used for the SDOF simulations
presented in the previous subsection. The analyses were conducted for harmonic excitations
with a duration of 10 seconds. The frequency of excitations of 5 Hz and 20 Hz were used. The
excitation was generated in a closed form, using harmonic functions built in the program. The
results were reduced to PTA vs R M D plots, which are presented in Figure 5.6. The frequency
95

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

of excitations is indicated at the upper-right corner of each plot. The results from 3DOF and
SDOF simulation are identified with legends on the right side of the plots.
The results show a very good match between the SDOF and the 3DOF models. The sliding of
the block was initiated at the same acceleration level of about 0.4 l g for both frequencies. This
is in agreement with assumption that the acceleration to initiate sliding does not depend on the
frequency of excitation for rigid body models. It can be concluded that both SDOF and 3DOF
numerical models predicted the same amount of sliding of the block.

5.1.3 Finite Element Model

5.1.3.1 Previous FE Modelling at UBC


At the same time when the further experiments were under way, a study to simulate the shake
table tests using a finite element model was conducted (Rudolf, 1998). The numerical model to
simulate the shake table tests was developed during this study using a commercial program
A N S Y S 5.3 Multiphysics/University (SAS IP, 1996). This model is called the original model
in this section. The original model was used to simulate some of the tests and it performed
satisfactorily especially at low frequencies of base excitation. The phenomena observed during
the tests at higher frequencies, such as combined sliding and rocking, could be simulated with
the original model but they did not manifest themselves under the same conditions as it was
during the tests. Therefore, the study (Rudolf, 1998) recommended several enhancements to the
FE model. These were done by the author and are described in this section.

96

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

5.1.3.2 ANSYS 5.3 Multiphysics/University


A N S Y S is a general finite element program with a long history of use for many applications. It
is commercially available and has an extensive library of elements and numerous solution
options described in detail in the program documentation. The A N S Y S modelling gives the
analyst wide control over the input, solution, and output. Data output and postprocessing can
be a crucial consideration in studies like this, when many simulations are performed and the
amount of generated data is large. The advantage of using A N S Y S is that this process can be
easily controlled by batch files which can run for several days performing series of analyses.
The output data can be partially post-processed and later used as input for other software. The
graphical interface allows the user to generate custom plots of the deformed shapes and stress
contours at any time step. This option is very valuable during the process of building and
debugging the model, which is described next.

5.1.3.3 Description of the Model


The experimental setup was modelled as a two-dimensional solid. The model is shown in
Figure 5.7. The dam monolith, the Upper Surface Plate (USP), and the shake table were
modelled using a simple plane element. The Lower Surface Plate (LSP) was integrated into the
shake table to form a single thick concrete plate, called the shake table here. The combined
flexibility of the physical shake table, LSP, and the clamping devices between the model and the
USP, was modelled through the contact stiffnesses of the contact elements.

The contacts

between the monolith and the LSP were modelled using a point-to-surface contact element. The
elements are described in the next subsection.

97

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

200

CONCRETE

|^

1800

Figure 5.7: Schematic of the F E Model

5.1.3.4 Plane Element


The bilinear plane element PLANE42 (SAS IP, 1996a) was used, with the plane-stress option.
The geometry of the element is defined by four nodes each with two translational degrees of
freedom. The element input further includes thickness and orthotropic material properties. The

98

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

input properties were obtained from the experimental model, which was described in Chapter 3.
A total of 64 plane elements were used to describe the geometry.

5.1.3.5 Contact Element


The Contact48 2-D Point-To-Surface Contact Element (SAS IP, 1996a) was used to model the
friction interface between the block and the LSP (shake table). It is capable of representing
contact and sliding of a point to surface in two dimensions. The element consists of three nodes:
the contact node and two nodes creating the target surface. In this application, the contact
function was assigned to the nodes on the contact projections of the block (heel & toe) and the
target function to the nodes on the top surface of the shake table.
The algorithm for contact and sliding can be summarized as follows:

Start from a no-contact position, the contact node approaches the target surface.

Contact is made when the contact node penetrates the target surface.

During the contact, the reaction-displacement relationship is governed by two


linear stiffnesses: one in the direction normal to the target surface and the other
tangential to the target surface. These stiffnesses are input by the analyst.

For subsequent solution, the sticking force is equal to the normal reaction force on
the surface times the static coefficient of friction.

If the tangential reaction exceeds the sticking force in the course of contact,
friction sliding is initiated with the friction force equal to the normal force times
the kinetic coefficient of friction.

As the contact node departs from the target surface, the reactions drop to zero and

99

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

a gap is developed.
The input parameters for the element included the normal and tangential stiffnesses entered,
respectively, as 20,000 kN/m and 800,000 kN/m. These values were calibrated to match the
natural frequencies of the model. The other parameters for the element were friction
coefficients, entered as those measured during the static tests. The model had two contact
elements on its toe and two on the heel. The shake table top surface included five target surfaces
on the downstream side and five on the upstream side. A total 20 of contact elements were used.

5.2 CALIBRATION OF THE FE MODEL


5.2.1 Modifications of the Original FE Model
The F E model for the experimental setup, developed by Rudolf (1998), had several known
deviations from the physical model. The author tried to remove as many of these as possible.
The modifications incorporated in the model included:

The shape of the F E model was modified so that it was closer to that of the
experimental setup.

The mass density of the model's material was slightly changed so that the model had
the same total mass as that measured during the tests.

The modulus of elasticity of the monolith was slightly updated to match the natural
frequencies of the bonded experimental model.

The number of elements of the model was decreased. The reason was that in this
study a large number of the analyses needed to be performed and the computational
time using the original mesh would not be acceptable. The number of the plane
elements decreased from 140 to 64 and contact elements from 42 to 20. After this
100

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

reduction a trial simulation was run with both the original and the reduced models
with a very good agreement of sliding.

The normal and tangential springs of the contact elements were updated to match the
natural frequencies of the unbonded experimental model.

Damping expressed by stiffness and mass proportional parameters was slightly


changed. This change was accepted based on the comparison of the model's
response during no-slip tests with the F E simulations. This comparison is described
in Appendix A and as a result of it the damping was changed from 5% of critical in
the fist two natural modes to 3.5% in the first and the third modes.

Other modifications, such as those to analysis procedures, are described here.

5.2.2 Modal Analysis


Modal analysis was performed in order to obtain natural frequencies and modes of the
experimental model and compare these with the measured values. In addition, this analysis was
one of the tools to calibrate the FE model. The results of this analysis were used to calibrate the
stiffness properties of the F E model to match the natural frequencies obtained from the impact
hammer testing described in Chapter 4.
Two variations of the experimental model approximated by finite elements were subjected to
modal analysis: 1) model bonded at the toe and the heel, and 2) unbonded model. The method
used to extract modal shapes and corresponding frequencies was the Subspace Iteration Method
described in SAS IP (1996b). In modal analysis, the contact elements acted as linear springs.
This means that the gap feature of Contact48 element was suppressed and classical problem of
linear free vibrations was solved to calculate natural frequencies of the unbonded setup.

101

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

The experimental and calculated natural frequencies of the bonded model are presented in Table
5.1. Two of the four calculated natural frequencies, for which a corresponding experimental
value was found, are in a very good agreement with the experimental values. The other two
correspond to the natural modes, which were not captured during the impact hammer tests.
Figures 5.8 to 5.11 show the F E model for the experimental setup bonded to the base and the
first three modes of vibration of the bonded setup. In order to clearly visualize the character of
every mode, the deflection shapes in the figures were exaggerated.
Table 5.1: Natural Frequencies of the Experimental Model Bonded to the Base
No.

Experimental
Frequency (Hz)

Frequency from
FE Model (Hz)

Character of Natural Mode

66

65.8

the 1st cantilever mode, see Figure 5.9

not identified

115.4

vertical bending, see Figure 5.10

174

173.5

the 2nd cantilever mode, see Figure 5.11

not identified

271.7

higher cantilever bending mode

Table 5.2: Natural Frequencies of the Unbonded Experimental Model


No.

Experimental
Frequency (Hz)

27-34

Frequency from
FE Model (Hz)

29.7

Character of Natural Mode


rocking of the model due to deflections in
the vertical springs combined with the 1 st
cantilever bending mode, see Figure 5.13
(springs not shown in the figure)
up/down stretching and shortening of the
vertical springs combined with the 1st
cantilever bending mode, see Figure 5.14
(springs not shown in the figure)

53-63

48.9

140

143.3

the 2nd cantilever bending mode,


see Figure 5.15

not identified

226.7

higher cantilever bending mode

102

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

Figure 5.9: First Natural Mode of Experimental Model Bonded to Base; f=65.8 Hz
103

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

CHAPTER

t
H

so
45

k^

at
SI
46

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

3H

X'

TU

rz

7i

76

7
61

62

63

64

56

57

56

59

52

53

54

SO

47

48

49

55

r * isi

5^

r^

Figure 5.12: F E Model for Unbonded Experimental Model of Dam Monolith

Figure 5.13: First Natural Mode of Unbonded Experimental Model; f=29.7 Hz


105

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

The modal analysis of the unbonded model was performed and its results are presented in Table
5.2. The measured frequencies varied with slightly from test to test. These variations were
attributed to nonlinear effects such as rocking or gapping present during the impact hammer
tests with the unbonded setup. During the modal analysis the vertical springs in the contact
elements were calibrated so that the first calculated natural frequency was very close to the
measured frequency.

The resulting agreement between the other calculated and measured

frequencies was found acceptable.

While the first and the third natural frequencies were

matched satisfactorily, the calculated frequency corresponding to the second mode of vibration
was lower than the measured one. However, the difference of about 15% was considered
acceptable because the mode has mostly vertical coordinates and the frequency is high.

5.3 RESULTS FROM SIMULATIONS OF SHAKE TABLE TESTS


The numerical models developed in the previous section were tested against the results from
selected shake table tests. This verification of the numerical models was limited to the tests with
harmonic and EQ2 (PSD shape from the Northridge earthquake record) base excitations. In
addition, only the tests with dominant frequencies 5, 10, 15, 20 and 25 Hz were used and only
five tests for each combination of the base excitation and dominant frequency were simulated.

5.3.1 Parameters of Simulations


The simulations with all three numerical models were performed using the parameters of the
experimental setup from the further testing program. The simulated hydrostatic force on the
model was obtained from the experimental data, which assured that any fluctuations of this force
during the tests were reproduced in the simulations. The friction coefficients were considered
107

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

with the values obtained prior to the test being simulated. Base motions were inputted to all
models in a form of shake table base accelerations measured during the test being simulated.
The FE simulations of the selected shake table tests with harmonic excitations performed by
Rudolf (1998) were done using a closed-form harmonic base displacements as driving signals
for the numerical model of the shake table. However, in the current FE simulations all the base
motions were inputted as measured shake table accelerations. The reason for switching from
closed-form to measured base motions was that the shake table did not reproduce exactly the
inputted closed-form driving signals during the tests. As a result of this, if the closed-form
signals were used, the simulated base motions would not be exactly the same as those which
excited the experimental setup during the tests. The reason for switching from the displacement
driven to acceleration driven numerical model was that the displacement records measured
during the tests at high frequencies (20 and 25 Hz) were not recorded properly by the
displacement sensor built in the shake table because of very small amplitudes of the tests at high
frequencies.
The base acceleration records inputted to the 3DOF simulations had to be resampled to 66% of
the sampling rate of those for FE and SDOF simulations. This was because of the limitation,
which the software Working Model puts on the number of data points imported during the
analysis. As a result of this, some peaks, especially at the records with higher dominant
frequencies, were lowered or skipped. This resulted in overall lower amounts of sliding from
3DOF simulations compared to those with SDOF model. Both models yielded the same
amounts of sliding when closed-form excitation was used, as shown in Figure 5.6. Therefore,
resampling of the input records was the true reason, why SDOF and 3DOF simulations did not
108

CHAPTER

ANALYTICAL

STUDY

OF THE DAM

MONOLITH

MODEL

yield almost identical results during simulations, which will be shown in Figures 5.16 and 5.17.

5.3.2 Comparison of Amounts of Sliding


The comparison of measured and simulated amounts of sliding is provided in Figure 5.16 and
5.17, respectively, for the tests with harmonic and synthetic earthquake (EQ2) excitations. Both
figures contain five plots, one for each dominant frequency of base excitations. The plots
contain four sets of displacement (sliding) to PTA relationships and legend at the lower-right
corner is common for all plots in both figures. These relationships were measured or calculated
only at five points in each plot and the lines between the points from the same source
(experiment, FE model, SDOF model or 3 DOF model) represent only trends reached by each
model. The next two paragraphs contain comments on comparison of simulation results with
experimental values.
5.3.2.1

Simulations

of Harmonic

Tests

All comments in this paragraph are related to Figure 5.16 and all simulated values are compared
with the experimental displacements (sliding). It can be observed from Figure 5.16a that the
measured response of the experimental setup at 5 Hz was very well simulated by all numerical
models. The situation at 10 Hz (Figure 5.16b) is different. Here, the FE model yielded lower
displacements for lower PGA's but larger for higher PGA's. The SDOF and 3DOF results are
lower than the experimental. A good agreement between the numerical and experimental results
can be observed for thefirstthree simulations at 15 Hz (Figure 5.16c). However, for two high
PGA's at this frequency the FE and experimental results are about three times larger than those
from SDOF and 3DOF simulations. The agreement of FE and experimental displacements is

109

CHAPTER

a) Dominant Frequency: 5 Hz

0.3

0.4

0.5
PTA (g)

0.3

0.4

ANALYTICAL

STUDY OF THE DAM MONOLITH MODEL

b) Dominant Frequency: 10 Hz

i r
0.4
0.5
PTA (g)
1

0.6

0.7

c) Dominant Frequency: 15 Hz

0.6

0.

d) Dominant Frequency: 20 Hz

0.5
PTA (g)

0.6

0.7

0.4

0.5
PTA (g)

e) Dominant Frequency: 25 Hz

Legend for all plots:

0.4

Experiment
FE Model

SDOF Model

3DOF Model

0.5
PTA (g)

Figure 5.16: Comparison of SlidingfromTests and Simulations for Harmonic Excitations


110

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

good at 20 Hz (Figure 5.16d), but the SDOF and 3DOF results are lower. However, the trends
are the same for all sets of results. The situation is different at 25 Hz (Figure 5.16e) where a
very good match between F E and experimental values can be observed. The sliding at first
increases with increasing PGA, but for higher PGA's the displacements do not grow. This was
caused by dominant in-plane rocking character of the experimental model's response. The
SDOF and 3DOF models could not capture this and, consequently, their results do not follow
the same trend as the first two.
5.3.2.2

Simulations

of Synthetic Earthquake

(EQ2) Tests

Similarly to the previous paragraph, all simulated results in this paragraph are compared with
the experimental displacements (sliding). A l l comments provided here are related to the
information presented in Figure 5.17 coming from the shake table tests and simulations with the
EQ2 base excitations. The agreement of all numerical results with those from experiments is
very good at 5 Hz (Figure 5.17a). Only the F E based displacements are somewhat larger than
the rest. The agreement of all results is very good at 10 Hz (Figure 5.17b) even though the
SDOF and 3DOF values are lower especially at higher PGA's. The results for 15 Hz (Figure
5.17c) and 20 Hz (Figure 5.17d) are very similar. For these, the experimental and F E
displacements match very well, but SDOF and 3DOF simulations resulted in approximately
50% sliding amounts compared to the first two. However, the trends among all results are
similar. The results for 25 Hz (Figure 5.17e) show a good agreement between finite elements
and experiments, but the SDOF and 3DOF do not exhibit even similar trends.
5.3.2.3

Summary of

Comparisons

Several conclusions can be drawn from the comparison of measured and calculated amounts of
111

CHAPTER

ANALYTICAL

a) Dominant Frequency: 5 Hz

0.6
15

1.0

1.2

PTA(g)

1.4

1.6

0.8

1.0

1.2

PTA (g)

1.4

1.6

OF THE DAM

MONOLITH

MODEL

b) Dominant Frequency: 10 Hz

1.8

c) Dominant Frequency: 15 Hz

0.6
,

-i
0.8

STUDY

0.6
1c

1.8

0.8

1.0 1.2 1.4


PTA (g)

1.6

1.

d) Dominant Frequency: 20 Hz

0.6

0.8

1.0 1.2 1.4


PTA (g)

1.6

1.

e) Dominant Frequency: 25 Hz

TO |
Legend for all plots:

Experiment

-0

FE Model

- Q

SDOF Model
3DOF Model

1.0 1.2
PTA (g)

1.4

Figure 5.17: Comparison of Sliding from Tests and Simulations for EQ2 Excitations
112

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

sliding of the experimental model during the shake table tests and their simulations. The
agreement of all numerical models with the experiment is generally good for frequencies from
5 to 10 Hz. At the frequencies 15 and 20 Hz, the FE model could capture well the behaviour of
the experimental setup, while the SDOF and 3DOF models yielded smaller amounts of sliding.
However, the trends between the results from the rigid models followed similar trends as those
of the experiments and FE simulations. Comparisons at 25 Hz showed that the SDOF and 3DOF
results did not follow similar trends.
The above generalisation of performance of the three numerical models permits the conclusion:

The response at the frequencies from 5 to 10 Hz, that is about the frequencies from
Group 1, which was defined in Chapter 4, can be simulated satisfactorily with all
numerical models.

The response at the frequencies of 15 an 20 Hz, that is about the frequencies from
Group 2, can be simulated using the F E model. The SDOF and 3DOF models can
capture only certain trends in the response.

The response at the frequency of 25 Hz, that is about Group 3, can be simulated
using the F E model only. The SDOF and 3DOF models do not simulate response
of the model satisfactorily.

It can be concluded that the SDOF and the F E models give similar answers as far as the
dominant frequency of the base excitation is no more than about 50% of the first natural
frequency of the unbonded model.

For greater excitation frequencies, the disagreement

increases and, in fact, Figure 4.19 shows that at 22.5 Hz and 25 Hz, coupled with larger PTA,

113

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

the experiment showed upstream movement. The SDOF model cannot predict this upstream
motion, although the F E could, as shown by Rudolf (1998).
Rudolf (1998) showed that the simulated behaviour during the tests at high frequencies and with
strong base excitations included significant rocking and upstream sliding of the model. The
response simulated by FE model was found to be in an acceptable agreement with the measured
and observed response during the tests. The FE model used in this thesis was a modification of
the original F E model developed by Rudolf. Due to this and since the results from the lower
acceleration level F E simulations presented here matched satisfactorily those from the tests it
was assumed that the F E model used here would capture well the rocking and upstream sliding
phenomena studied by Rudolf (1998).
It follows from the above that the F E model could simulate the response of the experimental
setup over wide range of excitation frequencies and PGA's. The range where the rigid models
can be used is not as wide and it is limited to the frequencies of base excitations below 15 Hz.
No significant improvement in the results was found if the 3DOF model, that is 3-DOF model
capable of rocking and jumping, was used compared to the SDOF model.
Using the same Pentium II 300 M H z personal computer, one simulation using the F E model
took approximately 2.5 hours, with 3DOF model about 6 minutes and with SDOF model about
5 seconds.

It can be concluded that the SDOF model can be satisfactorily used for the

simulations with the dominant frequencies of the base excitations below 15 Hz. Above these
frequencies, the F E model is recommended. The 3DOF did not prove any advantage against
either of the other two models. Therefore, only the F E and SDOF models will be used in the

114

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

next chapter to simulate response of a single monolith of a concrete gravity dam.


The 3DOF model, which was found useful during the design of the experimental setup, will not
be used for any further simulations. A n analysis to find out why the 3DOF model did not
simulate rocking of the experimental setup is described in the next section.

5.4 ANALYSIS OF SLIDING/ROCKING OF 3DOF MODEL


It was observed during the numerical simulations with the 3DOF and SDOF models that, in
contrast with the experiments, none of the numerical simulations yielded upstream sliding of the
model. In addition, the 3DOF numerical model, which was capable of simulating rocking, did
not show rocking of the model in any of the numerical simulations performed. In an attempt to
explain these observations, the following analysis was done.
The case of the model resting on the base, as shown in Figure 5.18, is considered. The forces
acting on the model include:
weight of the model W=mg, where g is acceleration due to gravity and m is mass
of the model;

simulated hydrostatic force F;

horizontal inertia force F ;

horizontal friction force acting along the foundation interface.

I N

A l l the dimensions of the model shown in Figure 5.18 and its mass are considered to be fixed.
However, because it was observed that the simulated hydrostatics force changed during all
shake table tests, this is given a value from the range 980 N to 1050 N . Also, the friction

115

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

Figure 5.18: Kinematics of the Dam Monolith Model


coefficient p of the foundation interface changed as different surfaces were used during the tests
and therefore the friction coefficient is considered in the range 0.63 to 0.78. The above two
ranges yield four possible combinations of these two input parameters:
1) p = 0.63 a n d F = 9 8 0 N , 2 ) u = 0.63 a n d F = 1050 N , 3) p = 0.78 and F = 980 N , and 4)
p = 0.78 and F= 1050 N .
Two cases will be analysed: a) the inertia force F is acting downstream and; b) the inertia force
in

F is acting upstream.
in

Case a): If the force F is acting downstream, the model can slide downstream or rock about B:
in

Sliding downstream is governed by the equilibrium equation:


116

CHAPTER

in

+ F = m\ig or ma +.F

1) a

(5.7)

F
= u.g--

= m\ig,

which leads to: a


where a

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

(5.8)

is the acceleration to initiate sliding. For combinations 1) to 4), a has the values:
s

= 0.42g ;2) a

= 0.41g ;3) a

= 0.57g \ 4) a

= 0.56g

Initiation of rocking about point B is governed by the equilibrium equation:


F y + Fh = mgx' or ma y + Fh = mgx'
in

(5.9)

mgx' - Fh
which leads to: a

(5.10)

' m y
where a

'

is the acceleration to initiate rocking about point B. For combinations 1) to 4), a

has the following values: 1) and 3) a

It follows from a comparison of a

= 1.17g ; 2) and 4)

and a

= 1.16g .

for all the four combinations that a

is always

significantly higher than a . This means that for any of the considered combinations the
s

downstream sliding is always initiated before rocking about point B could be initiated.
Case b): If the inertia force F is acting upstream, the model can slide upstream or rock about A .
in

Sliding upstream is governed by the equilibrium equation:


F -F
in

= m\ig or ma -F

= m\ig ,

(5.11)

F
which leads to: a

where a

1) a

= \ig+-

(5.12)

is the acceleration to initiate sliding. For combinations 1) to 4), a

= 0.84g ;2) a

= 0.85g ; 3) a

= 0.99g ; 4) a

has the values:

= l.OOg

Initiation of rocking about point A is governed by the equilibrium equation:

117

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

F y - Hh = mgx or ma y - Hh = mgx
in

which leads to:


where a

(5.13)

a
r

mgx + Fh
my

(5.14)

is the acceleration to initiate rocking about point A . For combinations 1) to 4), a

has the following values: 1) and 3)

a = 1.05g ; 2) and 4) a = 1.07g .

It follows from a comparison of a

and a

for all the four combinations that a

is always

higher than a , which means that sliding downstream will be initiated before rocking about
s

point B could be initiated. However, if one closely compares values for combination 4), which
correspond to a high value of friction coefficient, it is obvious that sliding upstream and rocking
about point A could be initiated because the acceleration to initiate rocking a = 1.07g
r

relatively close to that initiate sliding a

is

= 1.00g.

It follows from the above analysis that for the ideal case of a rigid body numerical model with
no imperfections, the rocking about A should not be initiated. However, in the experimental
analysis dealing with an imperfect and flexible model, or in FE analysis dealing with a flexible
model, one can conclude that rocking of the model can be initiated by a pulse of upstream inertia
force. The results of the F E simulations lead to the same conclusion.

5.5 SUMMARY
A total of three numerical models to simulate the shake table tests were developed in this
chapter. These included:
Single Degree Of Freedom model - SDOF model;
3DOF model using commercial software Working Model - 3DOF model; and
118

CHAPTER

ANALYTICAL STUDY OF THE DAM MONOLITH MODEL

model using finite element commercial software A N S Y S - F E model.

Parameters of the rigid models, SDOF and 3DOF, were specified according to the characteristics
of the experimental setup. Parameters of the F E model were specified similarly, but some
additional parameters of this model were obtained from calibration of this model using the
impact hammer and no-slip shake table tests.
The numerical models were used to simulate selected shake table tests including sliding. The
FE model, which is computationally demanding, simulated satisfactorily the majority of the
tests over a considered range of the dominant frequencies of base excitations. The SDOF and
3DOF models were successful in the lower range up to 15 Hz. They did not work satisfactorily
at higher frequencies. Performance of these two models was about the same and because the
SDOF model is not as demanding on computational effort as the 3DOF model, the SDOF model
can be recommended for use at the lower frequencies, below 15 Hz.
A l l the conclusions given above were related to the unbonded experimental setup with the first
natural frequency of about 30 Hz. The SDOF model performed satisfactorily for the frequencies
up to 15 Hz, which is a half of the first natural frequency. If an unbonded structure of the same
shape, but different dimensions, is to be analysed using the developed numerical models, it can
be expected that the SDOF model will yield good results for simulations with the dominant
frequencies below one half of the first natural frequency of the structure. This, however, should
be verified with some simulations using the F E model.

119

A N A L Y T I C A L STUDY OF
A F U L L - S C A L E DAM MONOLITH

A n analytical study to simulate the response of the experimental model of a concrete gravity
dam monolith was described in the previous chapter. Sliding of the experimental model,
unbonded at the base, preloaded by the simulated hydrostatic force, and subjected to base
excitations, was obtained using three numerical models. Out of these three, the F E and SDOF
models, were selected for another series of numerical simulations.
Concrete gravity dams are typically constructed in blocks separated by vertical contraction
joints, which may or may not be keyed or grouted. The design and analysis of straight concrete
gravity dams, except for those in narrow valleys, is traditionally performed by assuming that
each block responds independently. For this reason, understanding 2-D behaviour of individual
monoliths is usually considered relevant and 2-D plane-stress models are usually employed to
estimate safety for stability and safety. In addition, if non-linear phenomena are being studied,
2-D models are usually employed in order to reduce computational effort.

This chapter

describes results from a series of numerical simulations, performed on a single monolith of a


full-size 45 m concrete gravity dam.

6.1 OBJECTIVES AND SCOPE OF THE ANALYTICAL STUDY


ON A FULL-SCALE MODEL
The objectives of this chapter were:

to modify the existing F E and SDOF models to simulate behaviour of the fullscale concrete gravity dam monolith;

to obtain modal characteristics of the full-scale monolith modelled by FE;

to perform a number of FE and SDOF simulations to calculate the response of the


full-scale monolith with varying water level of the reservoir, type and peak ground
120

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

acceleration of the base excitation record;

to compare results of numerical simulations obtained in terms of the dam


monolith's sliding using the F E model with those using the SDOF model.

The analyses in this chapter were limited to those with the F E or SDOF models of a single
monolith of a gravity dam. The loads simulated on the monolith included gravity, inertia,
hydrostatic and hydrodynamic forces. Other loads were neglected.

The dam-foundation

interface plane was considered only with friction forces between the monolith and the
foundation. The applied base excitations were limited to those in a single upstream/downstream
horizontal direction. Other components of earthquake motions were not simulated.

6.2 DESCRIPTION OF THE STRUCTURE


The dam-water-foundation system can be considered with different level of complexity in
analytical studies. Generally, it can be said that the more time consuming the analysis method
is, the simpler model researchers tend to use. This mostly holds even i f the latest computers are
employed. During nonlinear analysis of a concrete gravity dam, the analysis of the entire dam
is often, in order to reduce computational effort, replaced by analyses of selected independent
monolith of the dam. Adopting this concept, boundary conditions and loads on the selected
monolith are somewhat different from those on the monolith built-in the dam, but this
approximation is accepted by many researchers and practitioners, for example Fenves and
Chavez (1995) or Tinawi et al. (1998).
A single monolith of a concrete gravity dam of a total height 45 m was selected for analysis
described in this chapter. The monolith had the same shape as the experimental model used in
this study and therefore, the width of the monolith in the upstream/downstream direction was
36 m. The monolith with a unit thickness of 1 m in the cross-canyon was studied. Concrete
with modulus of elasticity equal to 27 GPa, mass density of 2580 kg/m and Poisson ratio of
3

121

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE DAM MONOLITH

0.22 w a s c o n s i d e r e d as m a t e r i a l for the m o n o l i t h . T h e s e values were obtained f r o m m a t e r i a l


tests c o n d u c t e d o n one o f B C H y d r o d a m s o f s i m i l a r shape a n d size ( P o w e r t e c h L a b s , 1996).

6.3 m
\<

F i g u r e 6 . 1 : S c h e m a t i c o f the F u l l - S c a l e M o n o l i t h Structure

T h e f o u n d a t i o n interface w a s selected h o r i z o n t a l as c a n be seen i n F i g u r e 6 . 1 . T h e f o u n d a t i o n


r o c k w a s c o n s i d e r e d w i t h the m o d u l u s o f elasticity o f 15 G P a and P o i s s o n ratio o f 0.25.

The

h o r i z o n t a l forces b e t w e e n the d a m a n d the f o u n d a t i o n w e r e l i m i t e d to those f r o m f r i c t i o n w i t h


the static c o e f f i c i e n t [i

= 1.05 a n d the k i n e t i c c o e f f i c i e n t \i

= 1.00 . O t h e r effects i n the

d a m - f o u n d a t i o n interface p l a n e , s u c h as c o h e s i o n or i n t e r l o c k i n g , w e r e not c o n s i d e r e d .
b e d r o c k w a s c o n s i d e r e d w i t h o u t mass i n order to m o d e l o n l y its stiffness feature.

The

In other

w o r d s , the f o u n d a t i o n w a s considered to represent o n l y the support c o n d i t i o n s for the d a m , but


m a k e n o attempt at representing j o i n t d a m - f o u n d a t i o n m o d a l behaviour.

T h e reservoir w a s c o n s i d e r e d w i t h the m a x i m u m water l e v e l o f 9 6 % , that w a s 4 3 . 2 m , o f the


total height o f the m o n o l i t h . T h e reservoir effects were l i m i t e d to hydrostatic a n d h y d r o d y n a m i c
pressures o n l y . O t h e r effects, s u c h as u p l i f t force, were not c o n s i d e r e d .

122

CHAPTER

ANALYTICAL

STUDY OF A FULL-SCALE

DAM

MONOLITH

The hydrodynamic effects of the reservoir were modelled using the added mass approximation
(Okamoto, 1973). According to this theory, the hydrodynamic pressurep at a depths can be
calculated as:
p(y)
where

hi ^y

()
61

is horizontal acceleration, w is mass density of water, H is the total depth of the

reservoir andy is a coordinate from 0 to H, defined in such a way thaty = 0 at the water level
andy = //at the reservoir bottom. It follows from Eq. (6.1) that the formula for the added mass
HD

a t m

e depths is:
HD(y)

i^JHy

(-)
6

The approximation of hydrodynamic pressures used here assumes the water to be


incompressible. Such approach is simple and computationally efficient, but it does not fully
capture the behaviour of the water vibrating in interaction with the dam. It would be more
accurate to consider the water to be a compressible medium, but such an analysis would have to
be done in the frequency domain, since the hydrodynamic pressures from the compressible
liquid would be frequency dependent. This approach was used, for example, by Chavez and
Fenves (1996) who performed the analysis in time and frequency domains simultaneously.
Such analysis is very time consuming and it was considered to be beyond the scope of this study.
The approximation adopted here gives larger estimates of hydrodynamic pressures compared to
the case when water is considered compressible and, therefore, was on conservative side.
In all analyses described in this chapter, the properties of the dam-water-foundation system were
assumed constant except for the water level. It was the objective of this study to calculate the
response of the base excited dam monolith for varied water levels. These were considered from
60% to 100% of the full reservoir with a step of 5%. The water level of 60% produces a
hydrostatic force equal to 36% of the hydrostatic force of the full reservoir and therefore, this

123

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

range was considered sufficient.

6.3 DESCRIPTION OF BASE EXCITATIONS


A total of three ground acceleration records were selected for the analyses in this chapter. In
order to subject the dam monolith to a wide range of time histories. These included:

an Eastern North American type near source earthquake - the 330 degrees
component of the 1985 Nahanni Earthquake, North West Territories, Canada,
measured at Slide Mountain (Naeim and Anderson, 1996), P G A of 0.33g, denoted
Nahanni earthquake.

a far source subduction earthquake - the East-West degrees component of the 1985
Michoacan, Mexico Earthquake, measured at Villita station (Naeim and Anderson,
1996), P G A of 0.13g, denoted Mexico earthquake.

a Western North American type near source earthquake - the North-South


component of the 1994 Northridge, California Earthquake measured at the Sylmar
County Hospital parking lot (CSMIP, 1994), P G A 0.84g, denoted Northridge
earthquake.

The original processed acceleration, velocity and displacement records of the above earthquakes
are shown in the parts a to c of Figures 6.2 to 6.4. The time axis, identified in part c of each
figure is common for the parts a, b and c. At the bottom parts, the figures also contain
acceleration spectra plots for 5% damping. These plots indicate, which SDOF systems would
be affected the most by each earthquake. The Nahanni earthquake (Figure 6.2d), which is often
considered as a typical Eastern type earthquake motion, would affect the most systems with
natural periods of 0.5 second and also those with the natural period from 0.05 to 0.1 second. The
systems with natural period of 0.6 second (Figure 6.3d) would be affected the most by the
Mexico earthquake. The Northridge record would affect the most the SDOF systems with
124

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

natural periods between 0.3 and 0.4 second, as can be seen in Figure 6.4d.
Statistical distribution of the Peak Ground Acceleration (PGA) can be assumed to follow a
lognormal (Foschi, 1998) distribution and it obeys a relationship of the form (Madsen et al.,
1986):
PGA

= - ^ e ^ ' n O

+ F)
2

For an assumed mean value of this distribution a ^ O . l g and its coefficient of variation F=0.6,
a value of P G A can be calculated for a given value of the standard normal variable R . The
N

value of R is associated with the probability of the earthquake exceeding the peak ground
N

acceleration equal to PGA. Here, a total of four values of R equal to 3, 3.5, 4, and 4.5 were
N

considered and using these in Eq. (6.3) yielded the values of P G A equal to 0.453g, 0.597g,
0.788g and 1.040g. Such values of R were
N

used in order to cover a wide range of P G A ' s for

known exceedence probabilities. The spectral plots at the bottom parts of Figures 6.2 to 6.4
were calculated for all these four values of PGA.
It should be noted that in an analysis of sliding of an actual concrete gravity dam the uplift force
and the vertical base motions should be considered. It is recognised that both these loads would
have an adverse effect on the overall amount of horizontal sliding. The SDOF model developed
for this study was not able to take into account these types of loads. For simplicity, these were
not considered also in the finite element analysis in this chapter. The methodology of the
reliability study presented in the next section would remain the same should these two loads be
considered.

6.4 FINITE ELEMENT STUDY


6.4.1 Description of the FE Model for the Dam Monolith
The numerical model for the dam monolith was similar to that in the analytical study of the
125

CHAPTER

ANALYTICAL

STUDY OF A FULL-SCALE

DAM MONOLITH

a) Original Record - Ground Acceleration

10
Time (sec)
d) Amplified Records - Spectral Acceleration for 5% damping

1.5
Period (sec)
Figure 6.2: Characteristics of Selected Record from 1985 Nahanni Earthquake
126

CHAPTER

ANALYTICAL

STUDY

OF A FULL-SCALE

DAM

MONOLITH

a) Original Record - Ground Acceleration

20

Time (sec)

30

d) Amplified Records - Spectral Acceleration for 5% damping

c
g

ra 2
a
o
o

< .

1.5
Period (sec)
Figure 6.3: Characteristics of Selected Record from 1985 Mexico Earthquake
127

CHAPTER

1.0
r

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

a) Original Record - Ground Acceleration

D)

0.5

C
g

2
<B
<D
O

0.0

o -0.5

<

-1.0

1.0
10

o
o

0)

>

0.5
0.0
-0.5
-1.0
0.30

c) Original Record - Ground Displacement

0.15
c
E 0.00
0)
o
ai
a.
in - 0.15
b

0.30
10

Time (sec)
d) Amplified Records - Spectral Acceleration for 5% damping

1.5

Period (sec)
Figure 6.4: Characteristics of Selected Record from 1994 Northridge Earthquake

128

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

experimental model since the monolith was considered in two variations. These included the
models for the bonded and unbonded monolith.
Similarly to the F E model of the experimental setup, analysed in the previous chapter, the F E
model for the prototype dam monolith was designed to be relatively simple, with a small number
of finite elements, in order to keep the computational times of nonlinear time history analyses
at an acceptable level. The F E model of the monolith comprised, respectively, 28 and 30 plane
stress quadrilateral bilinear elements to simulate the monolith and the foundation rock. The
foundation rock was modelled 20 m below and the same distance beyond the toe and the heel of
the monolith. In addition, the model contained 7 contact elements distributed along the dam
base to simulate the foundation interface zone between the monolith and the foundation rock.
Features of the plane stress and the contact elements were described in the previous section and
the inputted parameters of these elements corresponded to the material characteristics of the
dam monolith and the foundation rock described earlier.
In addition to the plane-stress and the contact elements, the F E model contained 6 point mass
elements, which were used to account for the added lumped masses at wetted nodal points on
the upstream side of the dam monolith. The MASS21 (SAS IP, 1996a) point mass element was
used with the option of unidirectional mass in the horizontal upstream/downstream direction
only. The added masses corresponding to each wetted node were calculated using Eq. (6.2).
Damping characteristics used for the F E model of the full-scale monolith were similar to those
for the F E model of the experimental setup.

For the full-scale monolith, damping was

considered with the value of 2% of critical for the 1st and 3rd natural modes, which is a typical
value for concrete structures. It was recognised that damping can be an important parameter for
seismic response of a concrete gravity dam, but more complex representation of damping would
be beyond the scope of the study. Detailed information on the F E model can be obtained in
Appendix B where an example of the input A N S Y S file is given.

129

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

The F E model for the dam monolith was developed in two modifications:

the dam monolith bonded to the base, shown in Figure 6.5;

the unbonded dam monolith freely standing on the base, shown in Figure 6.9.

A total of three studies were conducted using the FE model. These included:

6.4.2

modal analysis of the monolith bonded to the base;

modal analysis of the unbonded monolith;

nonlinear time history analysis of the unbonded monolith.

Modal Characteristics of the Dam Monolith

The objective of the modal analyses was to obtain natural frequencies and mode shapes of the
dam monolith bonded to the base as well as the unbonded one. These were calculated using the
Subspace Iteration Method described in SAS IP (1996b)
The natural frequencies and the characteristics of the natural modes for the bonded monolith are
presented in Table 6.1. The frequencies are listed for the dam monolith with no water in the
reservoir and for that with a full reservoir. The natural frequencies of the monolith with the full
reservoir are lower than those for the case without water. This is an expected observation.
Characteristics of the obtained natural modes are presented in the right column of Table 6.1
where references to the corresponding figures showing the modes are also given. In order to
clearly visualize the character of every mode, the deflection shapes in the figures were
exaggerated.

The natural frequencies of a 3-D model of a similar concrete gravity dam

calculated by B C Hydro (1995) corresponding, respectively, to the first mode with the
frequency of 5.9 Hz and the second mode with the frequency of 15.2 Hz were 7.5 and 13.9 Hz.
This comparison indicates that the 2-D FE model used to simulate behaviour of the bonded dam
monolith had natural frequencies close to those of the other model for a similar structure and,
therefore the 2-D F E model was found acceptable for further analysis.

130

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

Table 6.1: Natural Frequencies of the Dam Monolith Bonded to the Base
No.

Frequency
(Hz) (no water)

Frequency (Hz)
(100% water)

Character of Natural Mode

6.9

5.9

1st cantilever bending mode, see Figure 6.6

16.9

15.2

2nd cantilever bending mode, see Figure 6.7

18.8

17.4

3rd cantilever bending mode, see Figure 6.8

34.1

29.3

higher cantilever bending mode

The natural frequencies of the unbonded dam monolith were calculated using the numerical
model for this variation. During the modal analysis and the contact elements acted as linear
springs. The natural frequencies of the unbonded model are presented in the second and third
columns of Table 6.2 for the dam monolith with no water and for that with full reservoir,
respectively. Characteristics of the natural modes are given in the right column of the table
where references to the corresponding modal plots are also shown.
Table 6.2: Natural Frequencies of the Unbonded Dam Monolith
No.

Frequency (Hz)
(no water)

4.3

9.3

Frequency (Hz)
(100% water)

3.7

9.2

Character of Natural Mode


rocking of the model due to deflections in the
vertical springs combined with the 1st cantilever
bending mode, see Figure 6.10
(springs not shown in the figure)
up/down stretching and shortening of the vertical
springs combined with the 1 st cantilever bending
mode, see Figure 6.11
(springs not shown in the figure)

16.2

14.2

the 2nd cantilever bending mode,


see Figure 6.12

27.9

25.5

higher cantilever bending mode

131

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

Figure 6.6: First Natural Mode of a Full-Scale Dam Monolith Bonded to Base; f=6.9 Hz
132

CHAPTER

ANALYTICAL

STUDY

OF A FULL-SCALE

DAM

MONOLITH

Figure 6.7: Second Natural Mode of a Full-Scale Dam Monolith Bonded to Base; f=16.9 Hz

Figure 6.8: Third Natural Mode of a Full-Scale Dam Monolith Bonded to Base; f=18.8 Hz
133

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

Figure 6.10: First Natural Mode of an Unbonded Full-Scale Dam Monolith; f=4.3 Hz
134

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE DAM MONOLITH

\\

\\

Figure 6 . 1 1 : Second Natural Mode of an Unbonded Full-Scale Dam Monolith;

f=9.3

Hz

135

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

It can be observed from Table 6.2 that the natural frequencies and the character of the natural
modes are quite different compared to the FE model of the bonded dam monolith. The first two
natural modes have strong rigid body character and can be characterised by deflections of
vertical springs in the contact elements. This can be observed from Figures 6.10 and 6.11.
The influence of added mass reduced values of all natural frequencies, but this reduction is small
for the second natural mode. This is because this mode has dominant vertical coordinates and
the added mass in the horizontal direction did not change significantly the ratio of the stiffness
to mass ratio for this mode.

6.4.3

Nonlinear Time History Analysis Using FE Model

The F E model of the full-scale unbonded dam monolith was used for nonlinear time history
analyses with the earthquake records defined earlier in this chapter. A total of 108 analyses were
performed, each taking about 3 hours of computational time of a Pentium II 330 M H z personal
computer. During these simulations, three parameters of the load on the dam monolith were
varied. These included the type (record) and P G A of the base excitations and the water level.
During the A N S Y S simulations, a transient analysis was specified with the Frontal Direct
Equation Solver. The Newmark implicit integration procedure was applied with an amplitude
decay factor of 0.005 (SAS IP, 1996b). It was found in the course of the study that satisfactory
results were obtained with the automatic time-stepping on (SAS IP, 1996c), with bounds on the
time step between 0.005 and 0.00005 seconds. The program automatically adjusted the time
step to suit the current state of the system. As expected, with an increasingly nonlinear
behaviour of the model, the time step became shorter, and the progress of the analysis slower.
The results from the analyses were stored at every 0.005 second.
The loading on the model was applied in load steps. During the first load step, with the inertia
effects turned off (SAS IP, 1996c) the gravity forces were applied. The second step, with inertia
136

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

effects still turned off, included application of hydrostatic pressure. This was superposed to the
state of the dam monolith after the first step. The effects of hydrostatic pressure were lumped
into hydrostatic forces applied at wetted nodes on the upstream side of the monolith. In the load
steps following the second, the earthquake loading in terms of the horizontal acceleration on
the dam-foundation system was applied on the monolith at the stage after the second load step.
A series of two trial simulations using each earthquake record was performed.

The two

simulations with the same earthquake record of equal P G A were carried out with the base
motions applied in opposite direction. This was done in order to find out if reversing the record
would change significantly the calculated amount of the base sliding. The differences in the
sliding response using the records in opposite directions did not exceed 15% for any earthquake.
Further analyses were carried out in the direction, for which larger sliding was obtained.
Detailed information about the A N S Y S commands used during the solution phase of the
analysis can be found in Appendix B. The results of the FE simulations are presented together
with those from the SDOF study in the last section of this chapter.

6.5 STUDY USING SDOF MODEL


6.5.1 Description of SDOF Model for the Dam Monolith
A SDOF model for the unbonded dam monolith was developed based on the theory presented
in the previous chapter. Parameters of the SDOF model were obtained from the dimensions and
the material characteristics of the dam monolith. The total mass of the block was calculated
from the volume of the monolith and the mass density of concrete as 2.46E6 kg, based on the
mass density of 2580 kg/m . The friction coefficients were used with the same values as those
3

for the F E simulations, static and kinetic 1.05 and 1.00, respectively.
The parameters of loads derived from the water level of the reservoir included hydrostatic and

137

CHAPTER

hydrodynamic forces.

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

The hydrostatic force was calculated as a lumped force from the

hydrostatic pressure distribution and it had values from 3.3E6 N for 60% water level to 9.1 E6
N for the 100% water level. The hydrodynamic pressures were simulated using the added mass
concept given by Eq. (6.2). This equation was used to calculate the total lumped added mass
M,
HD

which had values from 380E3 kg for 60% water level to 1,069E3 kg for 100% water level:
HD

6.5.2

fm {y)dy
0
HD

= f\wjH~ydy
0 o

= ^zwH
iz

(6.4)

Nonlinear Time History Analysis Using SDOF Model

The analyses using the SDOF model were performed with the earthquake records described
earlier in this chapter. The time step of the SDOF analyses was 0.001 second, but the results
were stored at every 0.005 second. A total of 108 SDOF analyses were performed, each of these
took about 8 seconds of computing time of a Pentium II330 M H z I B M compatible computer.

6.6

COMPARISON OF RESULTS FROM THE FE


AND SDOF SIMULATIONS

A large number of the F E and SDOF analyses were performed in this chapter. Results of these
are compared in this section in order to see i f the SDOF model can satisfactorily simulate the
response of the dam monolith. Both series of analyses yielded the amounts of sliding of the dam
monolith for varied earthquake types, PGA's and water levels. These results were summarised
in Figure 6.13 in terms of the monolith displacement (sliding) to water level relationships for
varied P G A ' s and in Figure 6.14 in terms of the monolith displacement (sliding) to P G A
relationships for varied water levels. The above means that both figures present the same set of
results but in a different way.
Both Figures 6.13 and 6.14 contain six parts a to f. These are organised in two columns and
three rows in each figure. The results from the SDOF model are presented in the first column

138

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE DAM MONOLITH

while those from F E are given in the second. Results from each of the earthquake records are
in one row and the name of the earthquake is identified above every plot. Every obtained result,
the amount of sliding of the dam monolith for a given combination of P G A and water is
indicated with a dot in the figures. To show trends in the results, the dots for the same P G A and
water level, respectively, are connected with solid lines in Figures 6.13 and 6.14.
In each plot in Figure 6.13, the lowest sliding was obtained for the lowest P G A of 0.453g while
the largest amplitude of 1.04g yielded the largest amounts of sliding, for a given water level.
Legend shown in part a of the figure also holds for the other parts. In each plot in Figure 6.14,
the lowest sliding for a given P G A was obtained for the lowest water level of 60% while the
maximum water level yielded the maximum sliding. The results for other water levels are
ordered between these for 60% and 100% water levels in steps of 5%. Legend for part a also
holds for the rest of the figure.
Observations from Figures 6.13 and 6.14 include the following:

Neither model yielded consistently larger amounts of sliding than the other.

The SDOF-based amounts of sliding exhibited greater dependence on the reservoir


effects than the FE model (compare Figure 6.13 a to b, c to d, e to f). The amounts
of sliding for low water level from the SDOF model were smaller than those from
the F E model. The agreement was better for higher water levels.

The SDOF model showed greater dependence on the P G A than the F E model.
This can be observed from comparison of Figure 6.14 a to b, c to d, e to f. The
S D O F model yielded smaller sliding than the F E model for low P G A ' s . The
agreement is better for higher PGA's.

The two observations above lead to a conclusion that the F E model initiated
sliding sooner for low water levels and low P G A ' s than the S D O F model.
However, the SDOF model, for high water levels and P G A ' s yielded amounts of

139

CHAPTER

a) SDOF Model - Nahanni Earthquake

menl

DAM MONOLITH

b) FE Model - Nahanni Earthquake

PGA 1.04g

5
E 4

ANALYTICAL STUDY OF A FULL-SCALE

PGA 0.79g

PGA 0.60g
PGA 0.45g

I2
Q.
b

-!
0

I
60

1 ,
!
70
80
90
Water level (%)
i

1
100

c) SDOF Model - Mexico Earthquake

i"

70
80
90
Water level (%)
d) FE Model - Mexico Earthquake

70
80
90
Water level (%)
1 5

e) SDOF Model - Northridge Earthquake

60

70
80
90
Water level (%)

100

15 f) FE Model - Northridge Earthquake

60

70
80
90
Water level (%)

100

Figure 6.13: Dam Monolith Displacements (Sliding) vs. Reservoir Water Level
140

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

a) SDOF Model - Nahanni Earthquake

2Q c) SDOF Model - Mexico Earthquake

0.4

0.6

0.8
PGA(g)

10

DAM MONOLITH

b) FE Model - Nahanni Earthquake

0.4

- Mexico Earthquake

0.6

0.8
PGA(g)

1.0

Figure 6.14: Dam Monolith Displacements (Sliding) vs. Peak Ground Acceleration

CHAPTER

ANALYTICAL STUDY OF A FULL-SCALE

DAM MONOLITH

sliding close to those from the F E model.


Another comparison of the SDOF and F E results was done for selected simulations in time
domain. Results of this comparison are shown in Appendix C where selected time histories
from 6 SDOF and 6 F E simulations were plotted. Two simulations for each earthquake record,
one with P G A of 1.040g and the other with 0.788g were selected for the comparison, all were
for the 100% water level. It follows from data shown in Appendix C that the SDOF and F E
friction foundation interfaces did not perform precisely the same way. This was expected due
to the different modelling techniques used for each model. However, the time histories of
sliding from both models were not found significantly different in any case.
A good agreement between obtained base sliding from the F E and SDOF models can be justified
using the results from the analytical study on the experimental model. It was observed during
this study that both models yielded similar sliding for the base excitation frequencies of up to
about 15 Hz, which corresponded to approximately to 50% of the first natural frequency of the
unbonded experimental setup (Table 4.3). In the case of the full-scale monolith, the first natural
frequency was 4.3 Hz and the dominant frequencies of the earthquake records used were from
1.6 to 1.9 Hz, which is less than 50% of the fundamental frequency of the monolith. If the
observation from the analyses on the experimental setup is extended to the full-scale case it can
be said that the F E and SDOF models could give similar answers also in this case.
It is difficult to make any general conclusion about the performance of the SDOF model without
detailed analysis of trends the F E and SDOF results show. The amounts of sliding the SDOF
model yielded were very close to those from the F E simulations for some combinations of
loading parameters. The agreement was not as good for the others. It is desirable to obtain a
simple mathematical relationships between the sliding, P G A and the water level.

Such

relationships obtained by interpolating the results could be used to observe general trends and
differences in performance of the FE and SDOF models. This will be done in the next chapter.
142

R E L I A B I L I T Y S T U D Y OF
A F U L L - S C A L E DAM MONOLITH

7.1 BASIC CONCEPTS OF RELIABILITY ANALYSIS


The reliability of an engineering system is the probability that it will perform as required in
given conditions within a specified period of time. For example, the reliability level may be
expressed as the probability that the maximum deflection of a structure will not exceed a given
value in the next 50 years given the load conditions for the site (Madsen et al., 1986).
In addition to the design parameters, the performance of the system is controlled by a set of
variables, some representing geometric, mechanical, and material characteristics of the structure
and other characterizing the external effects such as the load demands.
From the probabilistic point of view, the variables are regarded as random and have to be
described in probabilistic terms, including their distribution types, mean values, standard
deviations and other characteristics. However, those with a small degree of uncertainty can be
treated as deterministic parameters represented only by their nominal values. The probabilistic
description of a random variable may be achieved by a) experiments, providing statistics and an
estimate for the corresponding distribution or; b) engineering experience and judgement when
such statistical data are lacking. In addition, in order to reduce computational effort, a presensitivity analysis can be employed to determine the variables that should be treated as random
(Foschi etal., 1989).
The implementation of the reliability analysis is based on a description of the limit state of
interest by a performance function G(X) as follows:

143

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

G(X) = G(x ,x ,...,x )


x

DAM MONOLITH

(7.1)

where X = (x x ,...,x-^) is an N-dimensional vector of intervening random variables. Some


T

of these may affect the demand on the system, denoted D, while the others influence the system
capacity C to withstand the demand. By convention, the performance function G is written as:
G = C-D

(7.2)

The system will then fail i f the combination of the intervening random variables results in the
value G < 0. The corresponding probability of such an event (P(G < 0)) is called the probability
of failure. Conversely, the combination of the intervening variables resulting in G > 0 will make
the system survive and the corresponding probability (P(G > 0)) is called reliability. The
situation when the performance function G = 0, is called the limit state failure surface. To
calculate G, a simulation model describing the problem of interest is needed. For example, in a
deflection problem, G can be expressed as:
G = S -S(x x ,...,x )
Q

lt

(7.3)

where S is the allowable and S is the actual deflection of a structure. In our case, the allowable
0

and the maximum displacements will correspond to amounts of sliding of a gravity dam during
an earthquake. The actual displacement is a result of a deterministic calculation using a
numerical model. If this model can provide the required answer, is fast, and can be linked with
a reliability program, it can then be used directly in the reliability analysis. If getting the result
of one calculation takes a significant amount of time, it is convenient to pre-generate results and
reduce these into a form representing the response of the dam to varied input variables.
Using the response representation, the probability of failure can then be obtained by calculating

144

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

the probability of the event G < 0. As there could be a number of random variables involved in
G, the exact calculation could be obtained by integrating the joint probability density function
of all random variables over the failure region G < 0 (Madsen et al., 1986). This exact approach
can hardly be applied since the joint probability function is unknown and difficult to find.
Another possibility to calculate the probability of failure is standard computer simulation, the
Monte Carlo Method, which is simple to implement and can converge to the exact solution.
However, it could be very computationally demanding, especially when dealing with low
probabilities of failure. In such cases, the performance function has to be evaluated a very large
number of times to get a single outcome G < 0. In particular, i f each evaluation of the
performance function requires a nonlinear dynamic time-domain analysis to be performed, this
method can be very lengthy even if the latest computers are used (Li, 1999).
A second alternative is the use of approximate methods developed during the last three decades.
These include the First and Second Order Reliability Methods (FORM and SORM), which are
based on the calculation of the reliability index P . From this index, the probability of failure
Pyand the reliability P can be estimated approximately as follows:
R

P
F

* 1-

<D(-P), and

O(-P)

<P(P)

(7.4)
(7.5)

where <D is the standard normal probability distribution function.


The F O R M procedure was developed in the 1970's (Hasofer and Lind, 1974, Rackwitz and
Fiessler, 1978) as an extension of the work in the late 1960's (Cornell, 1969) and has been
modified and improved since. During the last twenty years, the method has been implemented
145

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

in a number of reliability analysis programs and used in many engineering applications. For an
exact estimate of the probability, the method requires the random variables to be normally
distributed and uncorrelated, and the function G to be linear. If some of the random variables
are correlated or non-normal, transformations are required before the procedure is implemented.
The reliability analysis package R E L A N was used for the analysis during this study. This
program has been developed at the Department of Civil Engineering of the University of British
Columbia (Foschi et al., 1997). The program can be used to solve a number of various
reliability problems employing either Monte Carlo or F O R M / S O R M methods.
Two numerical models were used in the analytical study of a full-scale gravity dam monolith to
obtain amounts of the monolith's sliding for varied loading parameters. These were the FE and
the SDOF models.

The SDOF model provides a fast solution to any combination of

deterministic input variables and it could be easily linked with a reliability analysis program, in
this case, R E L A N . However, the F E model needs a significant amount of time to calculate a
single result and, therefore, it is convenient to use a response surface approximation i f this
model is employed. Since it was the objective of this chapter to compare the performance of
both numerical models, the response surface method was used with both.
In the following sections, the reliability analyses conducted within this study will be described.
This will include several tasks such as reduction of results obtained from the last chapter,
identification of random variables, and several types of reliability analyses.

146

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

7.2 OBJECTIVES AND SCOPE OF RELIABILITY ANALYSIS


Two sets of results were obtained from the analytical study on a full-scale dam monolith, which
was described in the previous chapter. The results, obtained using the F E and SDOF models,
predicted how much the monolith would slide under three different types of time histories of
varying amplitudes and with different reservoir levels. In reality, a concrete gravity dam at a
certain site could be hit by various time histories, the characteristics of which would depend on
many factors, such as the distance of the dam from the epicentre of the earthquake and the
intervening geology. Therefore, it is desirable to reduce the information generated in the
previous chapter in order to predict how the dam would perform under a general base motion.
The objectives of this chapter is to obtain annual probabilities of failure of the full-scale dam
monolith using the F E and SDOF results calculated in the previous chapter. Probability of
failure is defined as a probability of exceeding specified amount of sliding. Another objective
of this chapter is to express results of the reliability study in terms of safety factors.
The methods to achieve these objectives are:

To average the results of the analytical study on a full-scale dam in such a way
that the response of the dam is related to an average earthquake and the variability
in the time histories is expressed by an error term. The average earthquake is
defined as such base excitation which would cause the response of the dam
monolith to be the mean obtained from all records used.

To reduce the results by expressing the influence of the reservoir's water level on
the amount of sliding in terms of a simple relationship.

To identify intervening random variables and to obtain their statistical parameters.

147

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

To carry out several reliability-based analyses of varied complexity making use of


the averaged and reduced results.

To express the results of the reliability-based studies in terms of factors of safety


against sliding.

The studies conducted in this chapter were limited to those on a single full-scale monolith of a
gravity dam, which was analysed in the previous chapter. It would be incorrect to extrapolate
the results of these studies to dams varying significantly from this structure, but it can be
assumed that the results are applicable to a concrete gravity dam monolith of similar properties.

7.3 REDUCTION OF THE ANALYTICAL RESULTS


The analytical results were obtained in the previous chapter as sets of maximum horizontal
displacement of the dam monolith relative to its base. These displacements, or amounts of
sliding, were calculated using the FE and SDOF models for:

three different earthquake records - one from each of the Nahanni, Mexico and
Northridge earthquakes;
four values of P G A - 0.45g, 0.60g, 0.79g, and 1.04g;

nine varied water levels of the reservoir - 60% to 100% with a step of 5%.

The reasons for selecting these parameters were given in the previous chapter. In order to reduce
the amount of the results the following analyses were performed:

Logarithms of the P G A values and corresponding amounts of sliding were calculated.


This was done because the P G A values are assumed to follow a lognormal distribution
(Li and Foschi, 1998), which means that their logarithms follow a normal distribution.
This conversion to normal variables makes a reliability analysis easier especially i f no
148

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

reliability software is available.


Pairs consisting of P G A and amount of sliding for all earthquakes records and PGA's,
considered for a given water level, were replaced by a linear approximation. This was
obtained by applying linear regression to all the pairs for a given water level. In an
obtained equation, the logarithm of sliding (denoted S) is expressed as a linear function
of the logarithm of the P G A (denoted a ):
G

In (5) = a + bln(a )

(7.6)

a and b are the coefficients calculated by linear regression.

These coefficients were

calculated for all water levels in Appendix D and E, respectively, for the results from the
FE and SDOF analyses. It can be concluded from observation of Figures DI and E l that
the linear interpolation used in Eq. (7.6) provided a good fit on the results.
The sets of coefficients a and b from Eq.

(7.6),

obtained for each considered water level

of the reservoir, were replaced by linear relationships in terms of this water level:
a = a, + a h , and

(7.7)

b = b +bh

(7.8)

In the above equations, h is the water level of the reservoir and the coefficients a\, a , b\,
2

and b were obtained using linear regression on the sets of coefficients a and b. The water
2

level was expressed in percent of the maximum level at the reservoir. After the double
linearisation, a S to a relationship was obtained:
G

ln(S) = a + a h + (b + b h)\n(a )
{

(7.9)

with one set of the coefficients a\, a , b , and b for the results from each one of the FE and
2

SDOF studies. This form of the interpolation the S to a relationships was selected because
G

149

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

it gives a good fit over the range of water levels from 60 to 100% of the full reservoir as
can be seen Figures D2, D3 and E2, E3. In addition, Eq. (7.9) represents a linear function
of l n ( a ) , which is a normal random variable because a is assumed lognormal.
G

A set of errors was calculated as difference between the results from the simulations and
the values calculated using Eq. (7.9).

Basic statistics, the mean value and the standard deviation, were calculated for the set of
the errors.

The same statistics were obtained for the set o f l n ( a ) values


G

corresponding to all calculated logarithms of sliding. In addition, a correlation


coefficient between these two sets was obtained.
A series of calculations described above was performed using a commercially available
equation editor Mathcad 8 Professional (Mathsoft, 1998).

Detailed information on these

calculations can be found in Appendices D and E, respectively, for the results from FE and
SDOF analyses.
The calculation of the logarithms of S and a , explained in the first of the above described steps,
G

could be carried out directly with the results of the FE study. However, the results of the SDOF
study did not allow such direct analysis because their number had to be increased first by the
results from analysis with high levels of PGA.

This was due to the fact that the SDOF

simulations at low P G A levels yielded almost zero displacements, which would make
calculation of the coefficients a ,a ,b
l

and b difficult. To do that, the amount of sliding equal


2

to 1 mm was selected as a point where the sliding of the dam was considered to be initiated. The
SDOF numerical model was used to determine the P G A values needed to cause such amount of
sliding. Then, several series of SDOF simulations were performed using all three earthquake

150

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

records with P G A ' s above the values needed to displace the dam monolith 1 mm.
After the data from analytical study was reduced three different reliability analyses were carried
out on each set of the results. Before the results of these analyses are presented it is necessary
to introduce basic concepts used in such analysis.

7.4 RELIABILITY STUDY WITH TWO RANDOM VARIABLES


This chapter of the thesis deals with the problem of evaluating the risk of exceeding a specified
amount of sliding for the studied monolith of a concrete gravity dam under an earthquake.
Although many uncertainties are involved in the problem, the selection of random variables for
the reliability study was based on the parameters varied during the analytical study performed
in the previous chapter. A total of three variables were varied in this study: 1) the type of used
earthquake record, 2) P G A of the base motions, and 3) the reservoir's water level. These
variables were assumed to affect the most the resulting amount of sliding of a gravity dam
monolith loaded by an earthquake and therefore their influence was studied in the previous
chapter. The reliability analysis performed in this section uses random variables based on the
first two of the three above parameters.
In addition to the above identified random variables, there would be a great deal of randomness
in the characteristics of the foundation rock and the foundation interface. Out of all these
parameters, it is assumed that the modulus of elasticity and the friction coefficient of the
foundation interface are the most important (Fenves and Chavez, 1996). However, these two
parameters were not varied during the deterministic analyses performed in the previous chapter
because the structural system was considered deterministic and only the loading was considered
151

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

variable. A sensitivity analysis will be performed in the end of this chapter to assess the
influence on the results of the foundation rock modulus of elasticity and the friction coefficient
of the foundation interface.

7.4.1

Peak Ground Acceleration as a Random Variable

The P G A is widely used as a design parameter in seismic analysis of all types of structures and
is one of the major factors controlling the response of a structural system during an earthquake.
A total of four P G A ' s associated with the horizontal motions applied at the base of the dam of
interest were used in the analytical part of this study. These acceleration levels were calculated
from the distribution of PGA, or a , which is assumed to be lognormal according to ground
G

motion attenuation laws. Therefore, a

obeys the following relationship of the form (Madsen

et al., 1986):

where a

= _^ /Win(i
e

^)

(7.10)

is the mean value of the distribution of the peak ground accelerations a

during an

event, V is its coefficient of variation and i ? is a standard normal variable. From typical
N

attenuation laws, the standard deviation of l n ( a ) , expressed as Vln(l + V ) , is assumed


2

here to be equal to 0.55, to which a value of

0.6 is associated. This value, widely used

in seismic engineering, is consistent with many soil attenuation relationships (Foschi, 1998).
The seismicity of the site and the statistics of the peak ground acceleration can be used to
calculate the peak acceleration of Maximum Design Earthquake (MDE) for the site. This peak
acceleration usually corresponds to certain Annual Exceedence Probability (AEP) and will be
called a

AEP

here. For buildings, the A E P of 1/475 (NBCC, 1995) is used, and such annual
152

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

risk also corresponds to an exceedence probability of 0.1 in a 50-year window. Conversely, the
mean value a , can be calculated from the probability P that the peak acceleration a
M

exceeds the acceleration a

AEP

during an earthquake, coupled with the assumption that the

earthquake occurrences follow a Poisson process.


The practice for dams, which are designed and analysed in accordance with the Dam Safety
Guidelines (Canadian Dam Association, 1999), is not the same as for buildings. For safety
evaluation of dams, the A E P to calculate the value of o:

AEP

is dependent on the consequence

classification of the dam in accordance with the Dam Safety Guidelines. The usual minimum
criteria for M D E varies from AEP of 1/100 to 1/1000 for low consequence dams, 1/1000 to 1/
10,000 for high consequence dams and 1/10,000 for very high consequence dams. Since the
consequence classification of the studied concrete gravity dam monolith is not known and the
AEP typically used for buildings is within the range for low consequence dams, all analyses
described in this section are related to AEP of 1/475. However, it will be shown later in this
section that the results can be easily related to an arbitrary desired AEP.
Employing the above explained concept and for AEP of 1/475, the relationship for P and a
E

can be written:
_L

i_ -vP (G>"AEp)
e

where v is the Poisson process mean arrival rate. Here, v =0.1, which means that the
earthquakes under consideration occur, on average, once every 10 years. For v = 0.1, P is
E

then P = 0.0211. Since, from the definition of the probability for a variable lognormally
E

distributed, P can be expressed as:


E

153

CHAPTER
\-0(R

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

(7.12)

the associated normal variable R is found to be R - 2.0315. After rewriting Eq. (7.10) in the
N

form:

(7.13)

aM

R Jln(\

+V)
1

the relationship between the mean value of P G A and the value of a

AEP

can be obtained. For

= O.lg the corresponding design P G A is ^AEP ~ 0.265g .

The relationship between o

AEP

and the A E P depends only on the seismicity parameters of a

site. That is, given the coefficient of variation V and the Poisson process mean arrival rate v ,
o

AEP

can be obtained using Eqs (7.11) to (7.13) with the condition that the mean PGA, a , be
M

constant. For example, for V=0.6, v=0.1 and a^j=0.\g, the 100 year return acceleration,
denoted here as #ioo> is 0.66 times that for the 475 year return period
1000 and 1/10,000, respectively, the values of # 1000
than

475. Since the parameters V, v , and a

a n

(#475).

For A E P of 1/

d ioooo are 1.17 and 1.80 times larger


a

are fixed for a site, the corresponding

AEP

acceleration can be given by the 475 year return or any other return since these are related as
discussed above. The corresponding ^AEP f other A E P can be calculated. Here, a site with
r

= O . l g , V-0.6 and v=0.1 is assumed.

It was observed during the numerical simulations with FE and SDOF models that the sliding of
the dam monolith was not initiated until the P G A of the earthquake reached a certain level. If
an approximation of the sliding of the dam monolith in terms of Eq. (7.9) is to be used it is
necessary to limit the range where Eq. (7.9) is applicable. Since this equation would yield nonzero sliding of the monolith even for very low P G A levels it was decided to use in the reliability

154

CHAPTER

RELIABILITY

STUDY

OF A FULL-SCALE

DAM

MONOLITH

analysis only such values of P G A which cause sliding greater than 1 mm. This value was
considered as a point where the sliding of the dam monolith was initiated. The PGA values
corresponding to this amount of sliding were calculated using Eq. (7.9) and are shown in Figure
7.1. It can be observed from this figure that the PGA's decrease with increasing water level for
both models. In addition, it can be said that the SDOF model needs higher P G A ' s to produce
sliding of 1 mm.
The above described phenomena was considered when parameters of the P G A as a random
variable were being determined. Therefore, it was decided that P G A in the reliability analysis
should be considered with a lower bound equal to the acceleration values shown in Figure 7.1.
The lower bounds for the P G A distribution had different values for different water levels, in
addition to differences resulting from the unequal performance of the F E and SDOF models.

60

70

80
90
Water level (%)

100

Figure 7.1: Values of P G A Necessary to Cause Sliding of 1 mm

155

CHAPTER

7A . 2

RELIABILITY

STUDY

OF A FULL-SCALE

DAM

MONOLITH

Effect of Different Earthquake Records Treated as


an Error Random Variable

The type of time history is another important parameter in seismic analysis. Three different
earthquake records were used during the numerical simulations and the differences among
responses due to these were shown in the previous chapter. It was observed that the amount of
sliding for the same P G A and water level varied significantly with the type of earthquake record.
In order to take this phenomena into account in the reliability analysis a vector of errors was
calculated for each of the F E and SDOF result sets. These errors were obtained as a difference
between either the F E or SDOF results and the corresponding interpolated values of sliding
given by Eq. (7.9). The details of error calculation are presented in Appendices D and E,
respectively, for F E and SDOF analyses. It was decided to consider the errors as a normal
variable with statistics calculated in the appendices. These included the mean errors, obtained
for the FE and SDOF data as numbers very close to zero and the standard deviations calculated
as 0.790 for the F E data and 0.791 for the SDOF data.
In addition, the correlation coefficient between the PGA's and the errors was calculated. The
correlation was found very small for both the F E and SDOF data. The correlation coefficients
for the F E and SDOF data were obtained, respectively, as -0.01 and 0.004, as can be seen in
Appendices D and E.
Moreover, it should be mentioned that a distribution other than normal could have been assumed
for the errors. Such distribution should be capable of including positive and negative error
values.

The normal and uniform distributions would be acceptable.

Here, the normal

distribution was used.

156

CHAPTER

RELIABILITY

STUDY

OF A FULL-SCALE

DAM

MONOLITH

Finally, it should be understood that instead of using three historical records, any number of
either historical or synthetic base acceleration records consistent with a given response spectra
or power spectrum could have been used for this study. As an approximation, it was assumed
that the three earthquakes used provide enough variability in the type of base excitations.

7.4.3 Simplified Reliability Analysis


The characteristics of the two random variables obtained in the previous subsection permit a
simple reliability analysis to estimate probability of failure of the system for a given allowable
displacement (sliding). This probability is associated with a constant reservoir level.

reliability program such as R E L A N should be used to perform the analysis and results of such
analysis will be presented in the next subsection. Here, another possibility of calculating this
probability without a reliability program, in closed-form, will be shown.
It was concluded in the previous subsection that the distribution of P G A as a random variable
should be considered with a lower bound. However, if this lower bound is not considered, it is
possible to obtain a closed-form solution to calculate the probability of failure. It is desirable to
obtain such a solution because it can be used as a check for a RELAN-based analysis. The
simplified analysis is based on the procedure explained below.
Let us assume that reliability of a system with two random variables is studied and the
performance function G is a linear combination of these variables. If the random variables are
normally distributed then the reliability of the system P can be calculated (Foschi, 1994):
Q

P =

(7.14)

157

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE DAM MONOLITH

where G is the mean value of the performance function and a

is its standard deviation. If

the performance function considered is:


G = \n(S ) - [a + a h + (b + b h)\n(a )
0

+ s]

(7.15)

where SQ is an allowable displacement, l n ( a ) and are normal random variables, and the
G

other parameters in this equation are known, then G can be calculated:


G = ln(S ) - [a + a h + (b + b h)\n(a )
0

+ i]

(7.16)

where l n ( a ) and s , respectively, are the means of the logarithms of P G A ' s and the errors.
G

The standard deviation c

c = (b
G

where

\ (a )
n

+ b h)o

^ d "

can be obtained from the equation:

ln{ac)

+ 2(6, + M)pcr l n ( f l o ) a e

(7-17)

are, respectively, the standard deviations of the logarithms of PGA's

and the errors, and p is the correlation coefficient between these two variables. A l l these
quantities were calculated for the FE and SDOF results, respectively, in Appendices D and E.
Knowing the reliability P, the event probability of failure P can be calculated using Eq. (7.4).
E

The desired value of the annual probability of failure can then be calculated from the Poisson
process assumption:
P

= l-e~

v / >

(7.18)

The above described procedure was used to calculate the annual probabilities of failure P

for

both the FE and SDOF data, for nine water levels from 60% to 100% with a step of 5%, and for
six levels of allowable displacement 1, 2, 5, 10, 15 and 20 cm. Sliding limits were selected
arbitrarily between 1 cm, which was considered as a significant initiation of sliding, and 20 cm,

158

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

which was considered as a significant amount of sliding. Significant sliding is expected to cause
disruption to the functioning o f the drainage system in the dam and i n the foundation, resulting
i n dangerous increases i n uplift and internal pore pressures. Significant relative sliding between
adjacent monoliths could also result i n damage to waterstops and pose a serious leakage hazard.
Analyses to determine consequences o f the dam monolith sliding by a given amount is beyond
the scope o f this study.

The results o f the simplified reliability analysis with two random variables are not shown here
because the R E L A N - b a s e d analysis, performed in the next subsection, should supersede them
by its accuracy. However, the results are presented i n Figures D 6 and E 6 and when compared
against those from the R E L A N - b a s e d analysis, they exhibit a good agreement.
The annual probabilities from the simplified analyses (Figures D 6 and E6) are slightly lower
than those calculated in the next section and presented in Figure 7.2a and b.

The differences

are caused by neglecting the lower bound for the peak ground acceleration i n the simplified
analyses. The differences are larger, but always smaller than one order, for the SDOF-based
probabilities than for those from F E . This is because the lower bound was higher for S D O F
than for the F E data as it was shown in Figure 7.1. However, in spite o f the differences, it can
be concluded that the simplified analysis with two random variables provided a good first
approximation to the true solution.

7.4.4 Results of Reliability Analysis with Two Random Variables


A simplified reliability analysis with two random variables was described i n the previous
subsection.

Although it can yield valuable results i f a reliability analysis program is not

159

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

available, it is preferable to use the program to obtain reliability of a structure with higher
accuracy. Results of such analysis are described in this subsection.
The performance function implemented into the program R E L A N had a form according to Eq.
(7.15). The two random variables, the PGA and the error from different time histories were used
with their statistics determined previously. Since one of the variables, the PGA, had its lower
bound given by the P G A to cause sliding of 1 mm, the total event probability of failure was
obtained as combined probability of:

The probability that the calculated sliding is greater than the allowable one,
assuming that the P G A is greater than its lower bound, and

The probability that the P G A is greater than its lower bound.

The above can be written as:


P

= (P(S>S )\(a >a ))P(a >a )


Q

G0

(7.19)

G0

where S is the amount of sliding calculated using Eq. (7.9), a

G0

is a lower bound for P G A

obtained from Figure 7.1, and S is a given maximum allowable sliding of the dam monolith.
0

The event probabilities P were calculated from results of R E L A N analyses using Eq. (7.19)
E

and the annual probabilities of failure P were obtained using Eq. (7.18).
A

The annual probabilities P , which are the final result of this section of the reliability study are
A

presented in Figure 7.2, a and b, respectively, for the FE and SDOF data. The legend shown in
part b of the figure is common for both a and b. The probabilities in both a and b follow very
similar trend, and also the values from F E data are close to those using results of SDOF
simulations. The SDOF-based probabilities show stronger dependence on the water level

160

CHAPTER

1E-4

fi

RELIABILITY

STUDY

OF A FULL-SCALE

DAM

MONOLITH

a) Annual Probabilities of Failure Obtained from Results of FE Simulations

70

1E-4

80
Water Level (%)

90

100

b) Annual Probabilities of Failure Obtained from Results of SDOF Simulations

1E-5

TO
LL

1 1E-6
o
Sliding 1 cm

<

Sliding 2 cm
1E-7

Sliding 5 cm
Sliding 10 cm
Sliding 15 cm
Sliding 20 cm

1E-8
60

70

80
Water Level (%)

90

100

Figure 7.2: Annual Probabilities of Failure vs. Water Level of the Reservoir for A E P of 1/475
161

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE DAM MONOLITH

compared to probabilities from F E data. On the other hand, the FE-based results are more
sensitive to the sliding criterion than those from SDOF analyses data. The SDOF probabilities
of failure are lower than those from F E for low water levels and low sliding. For example the
ratio of the FE to SDOF probability for 60% water level and 1 cm sliding is 2.4. Such ratios are
close to unity for water levels from 75% to 85% and sliding from 2 to 10 cm. It can be said that
in this range both models yielded almost the same results. The ratios are less than unity for high
water levels and large amounts of sliding. It can be concluded that the SDOF is unconservative
compared to the FE model for low water levels and small allowable sliding, but it is conservative
for high water levels and large allowable sliding. These observations are in a good agreement
with the conclusions about performance of both models, which were given in the end of the
previous chapter.
The reliability study with two random variables generated useful information. The probability
plots in Figure 7.2 can be used to determine the recommended water level of the reservoir i f a
target probability level and allowable sliding are given. For example, if the annual probability
of failure required is 1E-6 and the allowable sliding is 5 cm, the results using F E and SDOF
data indicate that the water level should be kept at 79% and 77%, respectively. For the same
probability level but allowable sliding increased to 10 cm, the FE-based results allow 93%
water level, while the SDOF-based only 87%. The plots can be also used to solve the inverse
problem, that is if a water level is given and the probabilities of failure associated with a given
allowable sliding need to be found.
The results obtained in this subsection allowed the calculation of the probability of failure i f a
water level is given. In practice, the problem to calculate a probability level for a range of water
162

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

level may need to be solved. This is because the water level could vary and it is not known what
it would be at the time of the earthquake. In such a case it is necessary to consider water level
as another random variable and modify the reliability analysis performed so far. This will be
done in the next section.

7.5

RELIABILITY STUDY WITH THREE RANDOM VARIABLES

In a more complete reliability analysis of sliding of a given concrete gravity dam, it is desirable
to obtain annual probabilities of failure if the dam is located in sites with different seismicity and
the water level of the reservoir fluctuates between given limits.
In the solution of such a problem, the acceleration
considered as a parameter. In this section a set of ^AEP
0.265g, 0.3g, 0.4g, 0.5g, and 0.6g. These

<*AEP

<*AEP
w

related to a given AEP, can be

considered with the values of 0.2g,

accelerations correspond to A E P of 1/475. As

it was explained earlier in this chapter, the corresponding peak accelerations for different AEP,
at the same site, can be obtained using Eqs (7.11) to (7.13). For each
the random variable a

QAEP>

distribution of

followed, again, a lognormal distribution with the coefficient of

variation V = 0.6. For v=0.1, the mean value a

varied for different values of

AEP

calculated according to Eq's. (7.11) to (7.13).


The water level of the reservoir was considered as a new random variable. It was assumed to
fluctuate between 60% and 100% of the full reservoir and it was uniformly distributed with a
mean value of 80%. If statistical data for water levels of a certain reservoir were available, some
other distribution fitting the data could be used. Since such data were not available, and it was
not the objective of this study to analyse any specific concrete gravity dam, the uniform
163

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

distribution was used.


The probability of exceeding a given sliding at a given site was calculated as in the previous
section with the exception that the variable h in the performance function G, see Eq. (7.15), was
now considered random. A series of RELAN-based analyses were carried out and the event
probability of failure P were calculated using Eq. (7.19). Having these, the annual probabilities
E

of failure P

were obtained using Eq. (7.18). Results of these calculations are presented in

graphical form in Figure 7.3. This figure shows the annual probabilities of failure P against
A

the mean peak acceleration a for the site and the corresponding peak acceleration a
M

AEP

for

two return periods of 475 and 10,000 years. The probabilities obtained from the F E and SDOF
results are shown, respectively, in parts a and b of the figure. The legend shown in part a of the
figure is common to both parts.
The results presented in Figure 7.3 show how the safety of a dam against base sliding would
change i f the dam is located in different sites. These results give the annual probability of
exceeding a given amount of sliding i f the site seismicity is given. This probability can be
obtained based on results either from F E or SDOF simulations. These results follow very
similar trends as can be observed from comparison of parts a and b of the figure. Similarly to
the previous section, considering allowable sliding as an independent parameter, the FE-based
probabilities are more sensitive to the sliding criterion than those from the SDOF data.
The results in Figure 7.3 correspond to A E P of 1/475 and 1/10,000. The relationship between
the corresponding values of a

AEP

was derived previously. The results in the figure can be

interpreted as follows: Using the F E based probabilities, the annual probability of exceeding a

164

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

a) Annual Probabilities of Failure Obtained from Results of FE Simulations

1E-2
1E-3
= 1E-4
o

5
co

"

Sliding 1 cm
Sliding 2 cm
Sliding 5 cm
Sliding 10 cm
Sliding 15 cm
Sliding 20 cm

o 1E-6

CL
"ro

1E-7
<
1E-8
1E-9
0.05
0.13
0.24

1E-2

0.10
0.26
0.48

0.15
0.40
0.71
Acceleration (g)

0.20
0.53
0.95

0.25
0.66
1.19

475
10000

b) Annual Probabilities of Failure Obtained from Results of SDOF Simulations

1E-3

'ro

1E-4
1E-5

15
ca

jQ

CL

1E-6

ro
3
C
C

<

1E-7

1E-8
1E-9
0.05
0.13
0.24

0.10
0.26
0.48

0.15
0.40
0.71
Acceleration (g)

"1
0.20
0.53
0.95

0.25 a
0.66

a475

1.19 a-ioooo

Figure 7.3: Annual Probabilities of Failure versus Acceleration


165

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

sliding of 10 cm equal to 1E-6 will be obtained in a site with approximately a^O.Wg


a =0.29g, or a =
475

]0000

or

0.52g.

It can observed from Figure 7.3 that the probabilities obtained from the SDOF-based results are
lower than those from the results of FE analyses for low peak acceleration and small allowable
sliding, but the differences are diminishing with increasing peak acceleration and allowable
sliding. For example, the FE-based probability of failure at the sliding level of 1 cm and

475

of 0.265g is almost 5 times larger than the corresponding probability yielded from SDOF data.
The differences in the probabilities are due to the different behaviour of both numerical models
and the lower values of SDOF-based probabilities are mainly due to the fact that the
accelerations to initiate sliding of 1 mm (see Figure 7.1) are higher than the peak accelerations
considered. Therefore, the combined probabilities according to Eq. (7.19) are lower for the
SDOF then for the F E data. Ratios of the probabilities from FE to those from SDOF are close
to unity from a

475

of OAg and sliding of 10 cm higher. In this range, the SDOF and FE models

yielded almost identical results.


A n example of results from the analysis with three random variables is presented in Table 7.1.
This table gives a summary of reliability analyses for a site with a

475

of OAg and for an

allowable sliding of 5 cm for both response surfaces, one obtained from the F E and the other
from the SDOF data. It can be observed from the table that the annual probability of failure
using the F E data was almost twice as that using the SDOF results.

166

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

Table 7.1: Example of Results for a site with a -0Ag


475

DAM MONOLITH

and for an Allowable Sliding of 5 cm

Reliability Analysis

Using FE Data

Using SDOF Data

Annual Probability
- of Failure

2.5E-5

1.3E-5

Random Variable

Design Point

Sensitivity Factor

Design Point

Sensitivity Factor

PGA (g)

0.79

0.84

0.95

0.81

Water Level (%)

89

0.30

91

0.44

Error due to Different


Earthquake Record

0.74

0.45

0.53

0.38

The results in Table 7.1 are given in terms of components of a design point and sensitivity
factors. The design point is defined as such combination of the input random variables for
which the probability of failure is the highest. It can be seen in the table that the components
of design point vary for the F E and SDOF based results. The design point peak ground
acceleration from the FE data is lower than the corresponding value from the analysis using the
SDOF model. In the contrary, the design point error due to different earthquake record is higher
for the FE results than that using the SDOF data. The design point water level is approximately
the same for both.
The sensitivity factors show how the resulting annual probability of failure depends on every
variable. It can be observed in Table 7.1 that the sensitivity factors for the peak ground
acceleration are about the same for the F E and SDOF data. However, they vary for both the
water level and the error due to different earthquake record. It can be concluded that for this site
with a

475

of OAg and an allowable sliding of 5 cm, the SDOF model is more sensitive to the

water level than the F E model, but it is the opposite for the error due to different time history.
Two sets of annual probabilities of failure for varied seismicity of the site and six levels of
167

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

allowable sliding were obtained in this subsection. Both sets yielded low annual probabilities
of failure from about 1E-8 to 1E-3, depending on the allowable sliding and the seismicity of the
dam site. The probabilities using the FE data are mostly higher, but not more than about 5 times
higher, than those from the SDOF data. In spite of these differences it can be concluded, that
results from both data sets are in a good agreement and the SDOF model performed
satisfactorily in this test.

7.6 RESULTS OF STUDY IN TERMS OF SAFETY FACTORS


The reliability analysis with three random variables was carried out in the previous section. One
of the intervening parameters was the seismicity of the site. This reliability analysis yielded
annual probabilities of exceeding a specified amount of sliding of the same concrete gravity dam
monolith at different sites. In this case, the acceleration corresponds to A E P of 1/475, but as
shown previously, the acceleration values could be scaled for any other A E P and the same set
of results could be used.
The concept of the factor of safety is a traditional procedure used in dam safety analysis. It is
desirable to express results from the previous section in terms of safety factors in simple design
or verification equations. The factor of safety is defined as a fraction between the nominal
capacity of a system and the nominal demand. In case of evaluating the stability of a gravity
dam monolith, shown if Figure 7.4, against sliding the factor of safety can be written as (Fenves
and Chavez, 1996):

r
F

Nominal Capacity

c
S

vi

Nominal Demand

^ t

T/

V Vt
+

168

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

S7_

Reservoir

Dam
Mo

Foundation Interface Plane

Foundation Rock
Figure 7.4: Forces on the interface zone of the gravity dam monolith
where N is the static normal force (resultant of self weight) and V is the static tangential
st

st

force (hydrostatic) on the interface zone AB, V is the maximum dynamic tangential force
(inertia and hydrodynamic). If the plane AB is not horizontal the forces have to be considered
with proper components. The properties of the interface zone are represented by the coefficient
of friction p and the total cohesion force C.
Several assumptions were used in the previous chapter during the development of numerical
models for the studied concrete gravity dam monolith. If these assumptions are taken into
account Eq. (7.20) can be rewritten in the following form:
(7.21)

where m = 2,456,000 kg is the mass of the monolith, H = 43.2 m is the maximum water level
of the reservoir, m

HD

_
L
y
^
_
^
_
^
/

is the added mass of water (Okamoto, 1973), h is the

relative water level expressed in% of H, g is the acceleration due to gravity, y is specific mass
169

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

of water, and the friction coefficient u, = 1.00.


The factor of safety can be calculated using the parameters from the previous section, that is for
the mean value for the water level h = 80% and for the six values of the acceleration a

475

Results of this calculation are shown in the first three columns of Table 7.2.
Table 7.2: Safety Factors Evaluated for Parameters Used in Reliability Study
Water Level (%)

Factor of Safety

P Sliding 1cm

P Sliding 10cm

0.2

80

2.0

2.6E-6

4.6E-8

0.265

80

1.7

1.7E-5

4.4E-7

0.3

80

1.6

3.6E-5

1.1E-6

0.4

80

1.3

1.9E-4

9.0E-6

0.5

80

1.1

6.0E-4

3.8E-5

0.6

80

1.0

1.4E-3

1.1E-4

<*475 (g)

The results presented in the third column of Table 7.2 can be linked to the results of the
reliability study with three random variables as follows. For example, the fourth row of Table
7.2, for a = OAg shows the factor of safety FS= 1.3. If this is compared with the results using
475

FE-based probabilities shown in Figure 7.3a, it can be concluded that this factor of safety
corresponds to the annual probability of exceeding sliding of 1 cm P = 1.9E-4, or to the annual
A

probability of exceeding sliding of 10 cm P = 9.0E-6. The values of P


A

of a

475

for the other values

were obtained from Figure 7.3a for the allowable sliding of 1 cm and 10 cm. These are

shown in the fourth and the fifth columns of Table 7.2.


The results presented in the first and the last rows of Table 7.2 can be interpreted as follows.
The studied dam, at the site with a =0.2g has the safety factor against base sliding of 2.0 and
475

the annual probability of sliding more than 1 cm equal to 2.6E-6. The same dam at a site with
170

7 RELIABILITY STUDY OF A FULL-SCALE DAM MON

CHAPTER

a =0.6g has the safety factor of 1.0 and the annual probability of sliding more than 1 cm equal
475

to 1.4E-3.
Another way in which the results of the reliability analysis can be used is to obtain a relationship
between the allowable sliding S and a
0

475

for a given target probability of failure. For example,

if the target annual probability of failure equal to 10E-5 is considered, the corresponding
acceleration a

475

can be obtained from Figure 7.3a for each allowable sliding. Then, for each

a , the corresponding factor of safety can be calculated using Eq. (7.21).


475

The above

described values are summarised in Table 7.3.


Table 7.3: Safety Factors Evaluated for the Target Probability of 1E-5 Based on A E P of 1/475
Allowable Sliding
So (cm)

Acceleration
475 (g)

Factor of Safety

0.25

1.79

0.28

1.66

0.35

1.44

10

0.41

1.31

15

0.45

1.23

20

0.48

1.17

The factors of safety given in the right column of Table 7.3 can be fitted with a quadratic
parabola as shown in Figure 7.5. The fitted values can be used in Eq. (7.21) instead of the safety
factors from Table 7.3:
A + BS + C S
0

2
0

^
a (m
D

HD

m)

(7.22)

.yg{ H)
m

where A = 1.816, B = -0.073 and C = 0.0021 are constants obtained by fitting the values of
safety factors with a least square technique. Eq. (7.22) gives the relationship in between S and
0

171

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

10
15
Allowable Sliding (cm)

Figure 7.5: Safety Factors for a Target Annual Probability of 1E-5 Based on A E P of 1/475
a

475

at a target probability level of IE-5. This relationship can be used to find a

475

(or S )
0

corresponding to S (or a ) if other values than those used in the previous analysis are desired.
0

475

Similar relationships could be also obtained for other probability levels.


When interpreting the information in Figure 7.5, one should keep in mind that the relationship
between the safety factor and the allowable sliding is for the same dam, for the constant
probability of failure (1E-5) and for the same A E P (1/475).

7.7 INFLUENCE OF FOUNDATION PROPERTIES


ON THE PROBABILITIES OF FAILURE
The probability of failure of a given gravity dam monolith subjected to base excitations was
studied in the previous sections. Loading parameters, such as the type of an earthquake, the
P G A of the base excitations and the water level of the reservoir were considered as random
variables in those analyses. The parameters of the system itself, such as the foundation rock
flexibility, coefficient of friction of the foundation interface, and others, were considered
172

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

deterministic and constant. In some situations it is desirable to analyse a system, which is not
exactly defined. In such a case its properties may vary at different locations of the system, and
they should be treated as random variables. It was shown (Fenves and Chavez, 1996) that the
amounts of sliding of a gravity dam are sensitive to variations in the friction coefficient of the
foundation interface and the foundation rock flexibility. Influence of these two parameters on
the amount of sliding of a dam monolith is studied in this subsection. The scope of this study
was limited to several simulations with the F E model, but the results of these simulations were
used to obtain annual probabilities of failure over a wide range of input parameters.
The modulus of elasticity of the rock foundation was considered with the value of 15 GPa in the
analytical study. For this value, for the Northridge earthquake record and the water level of
80%, the F E model yielded sliding of the dam monolith equal to 3.19 cm. When the modulus
of elasticity was changed to 10 and 20 GPa, the obtained sliding was 3.33 and 3.19 cm,
respectively. Therefore, it was concluded that the modulus of elasticity of the foundation rock,
in the tested range, does not have significant influence on the amount of sliding.
The situation was found different when the friction coefficient between the dam monolith and
the foundation rock was varied. Therefore, more analyses were performed and the results from
these are summarised in Table 7.5. A l l the analyses were performed with the Northridge
Earthquake record and with the water level of 80%. The P G A values of 0.60g, 0.78g and 1.04g,
were combined with the kinetic friction coefficients of 0.8,1.0 and 1.2.
In addition to the results in terms of sliding for varied friction coefficients, Table 7.5 gives ratios
of the amounts of sliding against those with p equal to 1.0. These are for p of 0.8 and 1.2,

173

CHAPTER

7 RELIABILITY STUDY OF A FULL-SCALE DAM MON

respectively, in the third and sixth columns of the table. Using these ratios, the sliding for any
obtained from the F E data, one using three and the other four random variables. It can be
observed from the table that the annual probability of failure from the analysis with four
random variables is about 1.3 times higher than that using three variables.
Table 7.4: Example of Results for a site with a =0Ag and for an Allowable Sliding of 5 cm
475

Reliability Analysis

with 3 Random Variables

with 4 Random Variables

Annual Probability
of Failure

2.5E-5

3.2E-5

Random Variable

Design Point

Sensitivity Factor

Design Point

Sensitivity Factor

PGA (g)

0.79

0.84

0.75

0.80

Water Level (%)

89

0.30

89

0.30

Error due to Different


Earthquake Record

0.74

0.45

0.70

0.44

Friction Coefficient

not applicable

not applicable

0.89

0.26

The results in Table 7.4 are given in terms of components of a design point and sensitivity
factors. It can be seen in the table that the components of the design point vary slightly for the
peak ground acceleration and the error due to the earthquake record, but are almost the same
for the water level. The sensitivity factors for the peak ground acceleration are the largest for
both cases, followed by the error due to different earthquake record and the water level. The
values of these three sensitivity factors are very similar for both types of analysis.

The

sensitivity factor for the friction coefficient is relatively small, which indicates that the
influence of the friction coefficient on the probability of failure is not as large as, for example,
that of the peak ground acceleration.
It can be concluded that for the variability considered in the friction coefficient, the annual
174

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE DAM MONOLITH

probabilities of sliding of a dam by a specified amount did not change significantly. It would
be an interesting study to calculate more sets of F E analyses, each for different friction
coefficient, over various earthquake records, PGA's and water levels, and use these results to
perform similar analyses as those in the previous subsections. However, time constraints on this
project do not permit such analyses.
Table 7.5: Results of FE Analyses with Varied Friction Coefficient
kinetic friction
coefficient
PGA (g)

7.8

p = 0.8

p= 1.0

p = 1.2

Sliding (cm) Ratio to Sliding Sliding (cm) Sliding (cm) Ratio to Sliding
for p = 1.0
for p=1.0

0.60

2.37

2.01

1.18

0.38

0.32

0.79

4.99

1.56

3.19

0.98

0.31

1.04

9.01

1.40

6.43

3.09

0.48

SUMMARY

It was the objective of this chapter to develop a procedure to perform a reliability analysis in"
order to find probabilities of sliding of a monolith of a full-scale gravity dam by a given amount.
It was also desired to assess if the probabilities obtained from the SDOF-based results are in a
good agreement with those from the results of FE analyses. First, the analytical response data
previously obtained during this study were reduced and linearised in order to capture general
trends in the data. Then, random variables for the reliability analyses were identified and a total
of three analyses were performed. Finally, a link between the results of one of the analyses and
safety factors against sliding was created.
The results of the reliability analysis with two random variables, presented in Figure 7.2,

175

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

indicate the overall low annual probabilities of sliding of the studied structure. These vary
significantly with the water level of the reservoir and with the allowable amount of sliding
considered. They range from about 7E-5 for 100% water level and 1 cm of allowable sliding to
2E-8 for the case of 60% water level and 20 cm of sliding allowed. Probabilities obtained from
the F E results are in a good agreement with those from the SDOF results.
The results of the study with three random variables yielded probabilities of failure i f the
acceleration corresponding to a given return period (AEP) and considered as one of the random
variables, was a parameter. The obtained probabilities from the F E and SDOF data were
presented in Figure 7.3. These results using both sources of data are in a good agreement even
though some differences could be noticed.
Using the FE data, the annual probabilities of failure range from 1.1E-8 for the a

475

acceleration

of 0.2g and 20 cm of allowable sliding to 1.3E-3 for 0.6g and 1 cm of sliding. If these results
are compared with those from the analysis with two random variables, a very good match is
obtained for both the F E and SDOF-based probabilities. This can be concluded if the range of
probabilities for water level of 80% from Figure 7.2a is compared with the range for the a

475

acceleration of 0.265g in Fig 7.3a. These ranges correspond to each other because the water
level of 80% was the mean value water level in the study with three variables and the value of
P G A equal to 0.265g was the a

475

acceleration considered in the analysis with two variables.

Similar conclusion can be made if the corresponding ranges of annual probabilities from Figure
7.2b are compared with those from Fig 7.3b.
The results of the reliability study were expressed in terms of safety factors. This provides a

176

CHAPTER

RELIABILITY STUDY OF A FULL-SCALE

DAM MONOLITH

link between the reliability study and the traditional concept used in dam safety analysis. It was
also shown, how the results of this study can be used to obtain relationship between the
allowable sliding and the seismicity of the dam's site for a given target probability and return
period of the earthquake.
In the end of the chapter, sensitivity of probabilities of failure against two parameters of the
foundation rock was studied. This study showed that small variations in the modulus of
elasticity of the foundation rock are not as significant as those in the friction coefficient. It was
shown, that if the friction coefficient is considered random with the mean being the same as the
deterministic value used previously, the annual probabilities of failure increased slightly but this
was not significant.
It was observed that the overall results in terms of various annual probabilities of sliding
obtained from the SDOF-based data do not vary significantly from those of the F E analyses.
Therefore, it can be concluded that the SDOF model is an adequate approximation for the
reliability analyses of concrete gravity dams of similar characteristics as the structure studied
here. This is a desirable conclusion, since the F E model, due to the longer computational time
needed for analysis it requires, makes the reliability analysis more demanding.

177

8 SUMMARY, CONCLUSIONS,
AND RECOMMENDATIONS
8.1 SUMMARY AND CONCLUSIONS
This thesis addressed the topic of base sliding of concrete gravity dams in experimental,
analytical and reliability studies. The background to the topic was discussed and the pertinent
research to-date reviewed. It can be stated that the objectives of all three studies were achieved
and the studies yielded ample results, observations and comments. Summary and conclusions
from the experiments, the analytical work and the reliability study are presented in the following
subsections.

8.1.1 Experimental Study


During the preliminary experimental study, a model of a single monolith of a concrete gravity
dam was designed and built. The components of the experimental model and setup were
described and the similitude aspects were discussed. The enhancements of the experimental
setup and testing procedures as a result of the preliminary tests were described.
Further experimental study of the model for the dam monolith included static and dynamic
shake table tests and impact hammer tests. A large amount of data was gathered during the tests.
These data were used for the verification and calibration of the numerical models during the
analytical study. Shake table testing manifested itself as a powerful tool to measure sliding of
base excited models of concrete gravity dams.
The natural frequencies of the experimental model obtained from the impact hammer tests were
found strongly dependent on the boundary conditions. The fundamental frequency of the
178

CHAPTER

SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS

unbonded model setup was determined to be in a range from 27 to 34 Hz while the one for the
bonded model setup was 66 Hz.
The static and kinetic friction coefficients of the interface between the model and its base were
extracted using the data from the static shake table tests. Both coefficients were found to
increase a little as the tests progressed
During the shake table dynamic tests, the response of the experimental unbonded model of a
gravity dam monolith was measured in terms of sliding of the model preloaded by a simulated
hydrostatic force and subjected to different types of base excitations of varying amplitude and
dominant frequency. The results showed that different records, either harmonic motions or
synthetic earthquakes, made the model respond in a similar way. The amplitude of base
acceleration and its dominant frequency were the parameters controlling the response.

In

Chapter 4, the dominant frequencies of the base excitations applied to the model were divided
into three groups, covering the testing range from 5 to 25 Hz. This grouping corresponded to
the changing character of the response of the experimental model in the shake table dynamic
tests. This response included downstream sliding (Group 1 - frequencies 5 to 12.5 Hz), sliding
combined with in-plane rocking (Group 2 - frequencies 15 to 20 Hz), and dominant rocking with
no sliding or sliding either downstream or upstream (Group 3 - frequencies 22.5 to 25 Hz).
A n upstream sliding consistently observed at high amplitudes of base accelerations with the
frequencies from Group 3 was concluded to be characteristic for the model response only.

8.1.2 Analytical Study


A total of three numerical models of varied complexity were developed during the analytical
179

CHAPTER

SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS

study of the scaled dam monolith model. These included SDOF, 3DOF and 2-D F E models.
The modal characteristics of the experimental setup obtained from the tests were used to
calibrate the FE model. The numerical models were developed in order to simulate the response
of the experimental setup preloaded by a simulated hydrostatic force and subjected to specified
base excitations.
The results using the SDOF, 3DOF and F E numerical models were compared with each other
and with the experimental data gathered during the shake table tests. The SDOF and 3DOF
models could simulate satisfactorily the response at low frequency the of base excitations, but
limitations of these models manifested themselves when the frequencies were increased. It was
observed, that the rigid models could simulate acceptably the sliding response of the
experimental model up to 15 Hz, which was about a half of its fundamental natural frequency.
The F E model could simulate satisfactorily the shake table tests throughout the entire range of
tested frequencies and amplitudes.
Analysis of combined sliding/rocking response of the 3DOF rigid numerical model was
performed during the analytical study. This analysis showed that for the considered geometry
of the rigid block and friction coefficient of the block-base interface, the base sliding was always
initiated before rocking. Therefore, no rocking was observed in the results from the simulations
using the 3DOF rigid model. The results of the 3DOF model using a commercial software were
similar to those from the SDOF model, but it was more convenient to use the latter for the
simulations during the analytical study on a full-scale monolith.
The SDOF and FE models were selected for the numerical analysis of a 45 m high monolith of

180

CHAPTER

SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS

a concrete gravity dam. The models were modified to describe the full-scale structure and to
capture the hydrodynamic effects of the reservoir. Modal characteristics of the bonded and
unbonded dam monolith were obtained using the F E model. The F E and SDOF models were
used to simulate the response of the dam monolith loaded by the hydrostatic and hydrodynamic
pressures and subjected to three historic earthquakes of varying PGA. The results using the FE
and SDOF models in terms of the base sliding of the monolith were compared.
In the analytical study of the full-scale monolith, a set of SDOF simulations yielded sliding
values close to those from the FE analyses. The sliding of the monolith using the FE model was
initiated at lower base accelerations than it was during SDOF simulations. However, the SDOF
model exhibited higher sensitivity to the PGA. A similar trend was observed for the reservoir's
water level. The results from the SDOF and F E models were close to each other for the
maximum PGA, which was about \g, and for high reservoir's water levels.

8.1.3 Reliability Study


The reliability study began with the reduction of data from the numerical simulations. This was
done to obtain the response surface describing base sliding of a concrete gravity dam monolith
preloaded by the reservoir's effects and subjected to specified base excitations. Then, random
variables were identified, described and their basic statistics were obtained. The random
variables, including the PGA, the error from the varying characteristics of the earthquake time
history, and the reservoir's water level were used in three reliability studies to determine annual
probabilities of failure. The failure was defined as exceeding a specified amount of base sliding
of the monolith during an earthquake.

181

CHAPTER

SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS

The annual probabilities of failure were obtained considering the three random variables. Using
the data from F E analyses, the annual probabilities of failure ranged from LIE-8 for the mean
P G A of 0.2g and 20 cm of the allowable sliding to 1.3E-3 for the mean P G A of 0.6g and 1 cm
of allowable sliding.
It was observed that the probabilities of sliding obtained from the SDOF data did not vary
significantly from those of the FE analyses. Therefore, it can be concluded that the SDOF model
is an adequate approximation for reliability analyses of concrete gravity dams of similar
characteristics as the structure studied here.
The results of the study were based on the annual exceedence probability (AEP) of 1/475, which
is a typical value used in building design. It was shown how the results can be used for safety
calculations of concrete gravity dams, in which different A E P ' s may be required depending on
the consequence classification of the dam.
In addition, the results of the reliability study were expressed in terms of the safety factors
against sliding. A relationship between the design acceleration and the allowable amount of
sliding was established for a selected target probability of failure. This provided a link between
the reliability study and the traditional concept used in the dam safety analysis. It was also
shown, how the results of this study can be used to obtain relationship between the allowable
sliding and the design acceleration for a given target probability.
The influence of randomness of the foundation rock on the amount of sliding was also studied.
This study showed that small variations in the modulus of elasticity of the foundation rock are
not as significant as those in the friction coefficient. It was shown, that if the friction coefficient

182

CHAPTER

SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS

is considered random with the mean being the same as the deterministic value used previously,
the annual probabilities of failure increased only slightly.

8.2 RECOMMENDATIONS FOR FURTHER RESEARCH


During this study, a large amount of valuable information was gathered. However, certain
simplifying assumptions had to be accepted in the course of the study and some of these could
be removed in further studies in order to obtain information applicable to a wider range of
problems. In addition, a few problems identified during the project were not addressed since
these were beyond the scope of the study. Recommendations for further research resulting from
these facts are summarised in this section.
During the experimental study, the sliding surfaces simulating the contact between the model
and its base, were made of concrete, which was the material of different properties than the rest
of the model. This was because the sliding surfaces were designed to last over many tests. After
an extensive material research, the sliding surfaces could be designed from some other material,
which would be of properties closer to those of the model's material and durable at the same
time. In addition, future experiments should address phenomena such as:

simulation of vertical earthquake motions,

simulation of bonding between the dam and its foundation,

simulation of hydrodynamic and uplift forces of the reservoir.

The FE model used in this study could be further developed. It is recommended in the A N S Y S
documentation (SAS IP, 1996c) that L S - D Y N A explicit integration program should be used for
highly geometrically nonlinear problems more exact results. This option should be investigated.

183

CHAPTER

SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS

It is the author's opinion that the solution to the classical mathematical equations (SDOF model)
would probably be the most promising research approach when other factors, such as vertical
ground motions and uplift, are to be considered. In addition, an oscillator could be used instead
of the rigid block to simulate effects of the fundamental mode of vibration.
Models other than Coulomb friction could be used during numerical simulations and these could
be calibrated using results of cyclic tests, which were conducted as additional experimental
study during this project.

Data from these tests are available at the Department of Civil

Engineering at U B C .
In current practice, uplift in cracks in a concrete gravity dam is typically assumed to remain
unchanged during the earthquake.

However, following the earthquake, uplift pressure in

earthquake-induced cracks would have time to build up. Post-earthquake safety is further
jeopardized should there be significant after-shocks with a relatively long time lag after the main
shock. Therefore, assessing post-earthquake safety of the dam is as important as assessing its
safety during the earthquake. Future research should consider including post-earthquake safety
in the numerical model.
A direct reliability analysis using the SDOF model linked with a reliability program could be
conducted and results of this could be compared with the current results using the response
surface method. Concept of inverse reliability could be applied to this problem, which would
permit to determine directly parameters of the system for a given target probability. In addition,
more earthquake records could be used in the reliability analysis to cover a wider range of
possible input motions.

184

REFERENCES
B C Hydro (1995). Ruskin Dam Deficiency Investigation. 3-D Finite Element Analysis and
Assessment of Dam Safety. Report No. MEP78, Dam Safety Deficiency Investigations,
Maintenance, Engineering and Projects Division, B C Hydro, Vancouver, B.C., Canada.
Bendat, J.S. and Piersol, A . G . (1971). Random Data: Analysis and Measurement Procedures.
John Wiley & Sons, Inc., New York, N Y , USA.
Black, C.J., T. Horyna, C E . Ventura & R.O. Foschi (1998). Shake table testing of a concrete
gravity dam model unbonded at the base. Research report No. 98-01. University of British
Columbia, Department of Civil Engineering, Vancouver, B.C., Canada, 1-112.
California Department of Conservation (1983). Processed Data from the Strong-Motion
Records of the Imperial Valley Earthquake of 15 October 1979. Special Publication 65,
California Department of Conservation, Division of Mines and Geology, Sacramento,
California, USA.
Canadian Dam Association (1999). Dam Safety Guidelines. Canadian Dam Association,
Ottawa, ON.
Chavez, J.W. and Fenves, G.L. (1994). EAGD-SLIDE: A Computer Program for the
Earthquake Analysis of Concrete Gravity Dams Including Base Sliding. Report No. U C B /
EERC-94/02, Department of Civil Engineering, University of California, Berkeley, California,
USA.
Chavez, J.W. and Fenves, G.L. (1995). Earthquake Response of Concrete Gravity Dams
Including Base Sliding. Journal of Structural Engineering, ASCE, Vol. 121, No. 5, pp. 865-875.
Chavez, J.W. and Fenves, G.L. (1996). Evaluation of Earthquake Induced Sliding of Gravity
Dams. Proceedings of the 11th World Conference on Earthquake Engineering, Acapulco,
Mexico.
Chen, H.C., Chang, C.H., Huang, L.S., L i , T.C., Yang, C.Y. and Wang, T.C. (1974). Earthquakes
Induced by Reservoir Impounding and their Effect on Hsinfenkiang Dam. Scienta Sinica, 17(2),
pp.239-272.
Chopra, A . K . (1995). Dynamics of Structures. Prentice Hall, Englewood Cliffs, NJ, USA.
Chopra, A . K . and Chakrabarti, P. (1972). The Earthquake Experience at Koyna Dam and
Stresses in Concrete Gravity Dams. Earthquake Engineering and Structural Dynamics, Vol. 1,
pp.416-465.
Chopra, A . K . and Zhang, L . (1991). Earthquake-Induced Base Sliding of Concrete Gravity
Dams. Journal of Structural Engineering, ASCE, Vol. 112, No. 12.
Cornell, C. A . (1969). Structural Safety Specifications Based on Second Moment Reliability
Analysis. Final Publication, International Association for Bridge and Structural Engineering
Symposium on Concept of Safety and Methods of Design. London, England.
CSMIP (1992). Strong-Motion Records from the Landers, California Earthquake of June
28,1992. Report No. OSMS 94-07, California Strong Motion Instrumentation Program,
Sacramento, California, USA.
CSMIP (1994). Strong-Motion Recordsfromthe Northridge, California Earthquake of January

185

REFERENCES

17,1994. Report No. OSMS 92-09, California Strong Motion Instrumentation Program,
Sacramento, California, USA.
Danay, A . and Adeghe, L . N . (1993). Seismic-induced Slip of Concrete Gravity Dams. Journal
of Structural Engineering, ASCE, V o l . 119, No. 1, pp. 108-129.

Darbre, G.R. and Wolf, J.P. (1988). Criterion of Stability and Implementation Issues of Hybrid
Frequency-Time Domain Procedure
for Non-Linear Dynamic Analysis. Earthquake
Engineering and Structural Dynamics, ASCE, V o l . 16, No. 4, pp. 569-581.

Donlon, W.P. and Hall, J.F. (1991). Shaking Table Study of Concrete Gravity Dam Monoliths.
Earthquake Engineering and Structural Dynamics, V o l . 20, pp. 769-786.

Fan, B . H . & J.J. Sled. (1992). Seismic Evaluation of Gravity Dams - Practical Aspects.
Proceedings of the 10th World Conference on Earthquake Engineering,

Madrid, 4645-4650.

Fenves, G.L. and Chavez, J.W. (1996). Evaluation of Earthquake Induced Sliding in Gravity
Dams. Proceedings

of the 11th World Conference on Earthquake Engineering,

Acapulco,

Mexico.
Fenves, G.L. and Chopra, A . K . (1985). Effects of Reservoir Bottom Absorption and DamWater-Foundation Rock Interaction on Frequency response Functions for Concrete Gravity
Dams. Earthquake Engineering and Structural Dynamics, V o l . 1 3 , N o . l , pp.13-31.

Fenves, G.L. and Chopra, A . K . (1985a). Reservoir Bottom Absorption Effects in Earthquake
Response of Concrete Gravity Dams. Journal ofStructural Engineering ASCE, Vol. 111, No. 8,
pp.545-562.
Fok, K . L . , Hall, J.F. and Chopra, A . K . (1986).

EACD-3D: A Computer Program for Three-

dimensional Earthquake Analysis of Concrete Dams. Report No. UCB/EERC-86/09. University

of California, Berkeley, California, USA.


Foschi, R.O., Folz, B., Yao, F. (1989). Reliability Based Design of Wood Structures. Structural
Research Series Report No. 34, Department of Civil Engineering, The University of British
Columbia, Vancouver, Canada.
Foschi, R.O. (1994). Reliability Analysis of Engineering Structures. Class Notes for Course
CIVL518, Department of Civil Engineering, The University of British Columbia, Vancouver,
Canada.
Foschi, R.O., Folz, B., Yao, F. and L i , H . (1997). RELAN: Reliability Analysis Program

User's

Manual. Department of Civil Engineering, The University of British Columbia, Vancouver,


Canada.
Foschi, R.O. (1998). Reliability-Based Design in Earthquake Engineering and Hysteresis
Modelling. Proceedings of the Third International

Conference on Computational

Stochastic

Mechanics, Santorini, Greece.


Gasparini, D . A . & E.H. Vanmarcke (1976). SIMQKE

- A Program for Artificial

Motion

Generation, User's Manual. Massachusetts Institute of Technology, Department of Civil


Engineering, Cambridge, Massachusetts, USA.

186

REFERENCES
Hall, J.F. (1988). The Dynamic and Earthquake Behaviour of Concrete Dams: Review of
Experimental Behaviour and Observational Evidence. Soil Dynamics
and
Earthquake
Engineering,

Vol. 7, No. 2, pp. 57-121.

Hasofer, A . M . and Lind, N . C . (1974). A n Exact and Invariant First Order Reliability Format.
Journal of Engineering
Mechanics Division, A S C E , 100(1), 111-121.
Horyna, T., Ventura, C. E., Foschi, R.O. (1997). Shake Table Testingofa Concrete Gravity Dam
Unbonded at the Base. Research report No. 96-06 prepared for the Dam Safely Program,
Maintenance Engineering and Projects Division of B C Hydro. Department of Civil
Engineering, The University of British Columbia, Vancouver, B.C., pp. 1-67.

Model

Horyna, T., C E . Ventura, R.O.Foschi & B.H.Fan (1998). Shake table studies of sliding of a
concrete gravity dam model. Proceedings of the 11th European Conference on Earthquake
Engineering.
Balkema, Rotterdam.
Indermaur, W., Brenner, R.P. and Arasteh, T. (1992). The effects of the 1990 Manjil Earthquake
on Sefi Rud Buttress Dam. Dam Engineering, Vol. 2, No. 4, pp. 275-305.
Kawamoto, J.D. (1983)
Frequency-Time

Domain

Solution
Approach.

of Nonlinear

Dynamic

Structural

Systems

by a

Hybrid

Ph.D. Dissertation, Massachusetts Institute of Technology,

Cambridge, Massachusetts, USA.


Knowledge Revolution (1996). Working Model 2D version 4.0 - User's Manual. Knowledge
Revolution, San Mateo, California
Leger, P and Katsouli, M . (1989). Seismic Stability of Concrete Gravity Dams.
and Structural Dynamics, Vol.18, pp. 889-902.

Earthquake

Engineering

L i , H . and Foschi, R.O. (1998). A n Inverse Reliability Method and Its Application.
Safety, Elsevier, 20(3), 257-270.

Structural

L i , H . (1999). An Inverse Reliability Method and its Applications in Engineering Design. Ph.D.
Thesis, Department of Civil Engineering, The University of British Columbia, Vancouver,
Canada.
Lin, G., Zhou, J., Fan, C. (1993). Dynamic Model Rupture Test and Safety Evaluation of
Concrete Gravity Dams. Dam Engineering, Vol. IV, No. 3, pp. 769-786.
Madsen, H.O., Krenk, S. and Lind, N.C. (1986). Methods of Structural
Inc., Englewood Cliffs, NJ, USA.
Mathsoft (1998). Mathcad8

Professional-User's

Manual.

Safety.

Prentice Hall

Mathsoft, Inc. Cambridge, M A , USA.

Mir, R.A. and Taylor, C.A. (1995). A n Experimental Investigation into Earthquake-induced
Failure of Medium to Low Height Concrete Gravity Dams. Journal of Earthquake
Engineering
and Structural

Dynamics,

V o l . 24, p. 373-393.

Mir, R.A. and Taylor, C.A. (1996). A n Investigation Into the Base Sliding Response of Rigid
Concrete Gravity Dams to Dynamic Loading. Journal of Earthquake
Engineering
and
Structural Dynamics, Vol. 25, pp. 79-98.

187

REFERENCES

Mlakar, P.F. (1987). Non-linear Response of Concrete Gravity Dams to Strong EarthquakeInduced Ground Motions. Computers and Structures, Vol. 26, pp. 165-173.
Moncarz, P.D. and Krawinkler, H . (1981). Theory and Application

of Experimental

Model

Analysis in Earthquake Engineering. Report No. 50, The John A . Blume Earthquake
Engineering Centre, Stanford University, Palo Alto, C A , USA.
Naeim, F. and Anderson, J.C. (1996). Design Classification

of Horizontal

and Vertical

Earthquake Ground Motion (1933-1994). A Report to the USGS, John A. Martin & Associates,
Inc., Los Angeles, California, USA.
National Research Council (1990). Earthquake Engineering for Concrete dams:

Design,

Performance and Research Needs. National Research Council, Commission on Engineering


and Technical Systems, Committee on Earthquake Engineering, Washington, D.C., USA.
N B C C (1995). National Building Code of Canada. National Research Council of Canada,
Ottawa, O N .
Niwa, A . and Clough, R. W. (1980). Shaking Table Research on Concrete Gravity Dam Models.

Report No. U C B / E E R C 80-05, Earthquake Engineering Research Center, University of


California, Berkeley, California, USA.
Okamoto, S. (1973). Introduction to Earthquake Engineering. Wiley & Sons, New York, USA.
Powertech Labs Inc. (1996). Testing Concrete Cores from Ruskin Dam. Project #7978-32,
Powertech Labs Inc., Surrey, B.C., Canada.
Rackwitz, R. and Fiessler, B . (1978). Structural Reliability Under Combined Random Load
Sequences. Computers and Structures, 9, 484-494.
Reich, R.W., Cervenka, J. and Saouma, V . E . (1991). Numerical Techniques for 2D and 3D
Nonlinear Fracture Mechanics Based Analysis of Dams. Proceedings of International
Conference on Dam Fracture, pp. 163-184, Boulder, Colorado, USA.
Rezai, M . (1999). Seismic Behavior of Steel Plate Shear Walls by Shake Table Testing. Ph.D.

Thesis, Department of Civil Engineering, The University of British Columbia, Vancouver,


Canada.
Rudolf, J. (1998). Finite element verification of shake table testing of a gravity dam model.

Research report No. 98-02. University of British Columbia, Department of Civil Engineering,
Vancouver, B.C., Canada, 1-60.
SAS IP, Inc. (1996). ANSYS Finite Element Computer Program, Version 5.3, Swanson Analysis

Systems, Inc., Houston, Pennsylvania, USA.


SAS IP, Inc. (1996a). ANSYS Element Reference, Eighth Edition. Swanson Analysis Systems,

Inc., Houston, Pennsylvania, USA.


SAS IP, Inc. (1996b). ANSYS Theory Reference, Seventh Edition. Swanson Analysis Systems,

Inc., Houston, Pennsylvania, U S A .


SAS IP, Inc. (1996c). ANSYS Advanced Analysis Techniques Guide, First Edition. Swanson

Analysis Systems, Inc., Houston, Pennsylvania, USA.


188

REFERENCES

Simic, M . and Taylor, C A . (1995). Limited Tension Non-linearities of Concrete Gravity DamFoundation Systems Under Static Loading. Dam Engineering, Vol. 6, pp. 23-62.
Simic, M . and Taylor, C A . (1996). Analytical Studies of the Non-linear Behaviour of Concrete
Gravity Dam -Foundation Systems Under Earthquake Loading. Journal of Earthquake
Engineering and Structural Dynamics.
Stewart, J.P., Bray, J.D., Seed, R.B. and Sitar, N . (1994). Preliminary Report on the Principal
Geotechnical Aspects of the January 17, 1994 Northridge Earthquake. Earthquake Engineering
Research Center, Report No. UCB/EERC-94-08.
Taylor, C A . (1996). Non-Linear Analysis of the Seismic Behaviour of Moderate Height
Concrete Gravity Dams. Proceedings Of the 11th World Conference on Earthquake
Engineering, Acapulco, Mexico.
Tinawi, R., Leger, P., Ghrib, F., Bhattacharjee, S. and Leclerc, M . (1998a). Structural Safety of
Existing Concrete Dams: Influence of Construction Joints. Volume A. Review of Literature and
Background Material. Department of Civil Engineering, Ecole Polytechnique, Montreal,
Quebec, Canada. (CEA Project No.9032 G 905)
Tinawi, R., Leger, P., Ghrib, F., Bhattacharjee, S. and Leclerc, M . (1998b). Structural Safety of
Existing Concrete Dams: Influence of Construction Joints. Final Report. Department of Civil
Engineering, Ecole Polytechnique, Montreal, Quebec, Canada. (CEA Project No.9032 G 905)
Tinawi, R., Leger, P., Ghrib, F., Bhattacharjee, S. and Leclerc, M . (1998c). Structural Safety of
Existing Concrete Dams: Influence of Construction Joints. Volume B. Theoretical and
Numerical Developments and Case Studies. Department of Civil Engineering, Ecole
Polytechnique, Montreal, Quebec, Canada. (CEA Project No.9032 G 905)
Westermo, B . and Udwadia, F. (1983). Periodic Response of a Sliding Oscillator System to
Harmonic Excitatioa Earthquake Engineering and Structural Dynamics, Vol. 11, pp. 135-146.

189

APPENDIX A - PLOTS OF M O D E L MOTIONS FROM


N O - S L I P TESTS AND SIMULATIONS
A finite element numerical model to simulate response of the experimental model due to shake
table motions was developed in Chapter 5. In this chapter, the experimental model unbonded
at the base and preloaded by a simulated hydrostatic force was created using commercial
package A N S Y S (Swanson Analysis Systems, Inc. 1996). During the experimental part of the
study described in Chapter 4, several shake table tests with low level excitations were
conducted. The objective of these tests was to generate data for verification of the FE model for
response with no sliding of the model. Such verification is described in this appendix.
Figures A . l to A.9 show selected motions of the experimental model measured during the shake
table tests and those calculated by F E simulations of the same tests. These motions include for
each test:

Horizontal acceleration of the model relative to the shake table measured or calculated at
the top of the upstream side of the model. This is shown in part a of the figures.
Horizontal acceleration of the model relative to the shake table measured or calculated at
the base on the upstream side of the model. This is plotted in part b of the figures and the
legend shown in this part also holds for part a. The plotted experimental acceleration was
reduced by a fraction of the measured acceleration at the top of the model (part a) in
order to take into account that the sensor was not located exactly at the base of the dam,
but it was attached 18 cm above the plane through foundation interface.

Shake table horizontal acceleration plotted in part c.

Shake table horizontal displacement plotted in part d.

Each of the Figures A . l to A.9 presents a 1 second long close-up on results from one test with
harmonic base excitations. The amplitudes of shake table motions for all tests were about 0.3g,
but the frequency of excitation varied for each tests ranging from 5 Hz to 25 Hz as can be seen
in captions at the figures. The observations made from the figures follow.
Shapes of the shake table acceleration plots are close but not exactly the same as those of a
closed-form harmonic excitation. This distortion is due to the interaction of the shake table with
the model during the tests. Since this distortion may be even amplified for tests with high level
of excitations, it is necessary to use acceleration records measured during the tests for the
numerical simulations of the tests. Using the closed-form base excitations as it was done in the
previous study (Rudolf 1998) brings uncertainty to the comparison with the experimental
results, which can be eliminated by using measured acceleration records.
The plots of experimental relative accelerations at the base of the monolith, presented in part b
of Figures A . l to A.9, show that some differential motions with respect to the shake table were
measured during no-slip tests in all frequencies. This was probably caused by the fact that the
upper surface plate was attached to the experimental model only at the heel and toe of its base,
190

APPENDIX A - PLOTS OF MODEL MOTIONS FROM NO-SUP

TESTS AND SIMULATIONS

which permitted differential accelerations with amplitudes ranging from 0.05g to O.lg. The
observed differential motions are small and most likely did not affect significantly the sliding of
the experimental model during the shake table tests including sliding of the model. It can be
also observed from the plots that similar response was not found in results from the numerical
model. This was due to the fact that in the FE model the upper surface plate was rigidly attached
to the monolith along the entire base, which did not allow any differential motions such as those
observed in the experimental results. The upper surface plate of the FE model could be attached
similarly as it was in the experimental model, but doing this would have required to generate a
number of new contact elements between the monolith and the plate and increase the
computational effort. Considering this and the fact that the upper surface plate has no meaning
for the case of a full-scale concrete gravity dam, the plate was modelled firmly attached to the
monolith in the FE study.
Comparisons of measured and calculated relative accelerations at the top of the monolith show
that these are generally in a good agreement. This conclusion can be made from visual
observation of the time histories shown in Figures A.l to A.9, parts a and b. The amplitudes of
the measured accelerations are in some cases larger than those calculated during simulations, in
other cases in simulations, and in some cases they are about the same. It can be said that the
amplitudes are generally in a good agreement.
A good agreement between experimental and analytical results from no-slip tests was reached
by decreasing the damping of the FE model compared to previous work of Rudolf (1998).
Damping in ANSYS model can be approximated using the concept of Rayleigh damping
(Chopra 1995) involving stiffness and mass proportional parameters ct and P. In the previous
work, these parameters were specified based on the time history results of impact hammer tests
on a full experimental setup. The decay of amplitudes of horizontal acceleration at the top of the
monolith was measured. Combined with the known natural frequencies, the calculated cc and
P were 8.0 and 0.0003, respectively, resulting in a damping ratio of approximately 5% in the
first two modes. In the current work, the damping parameters a and P were considered 11.5
and 0.000065, which corresponded to 3.5% of critical damping for the first and third modes.
These values of cc and P are very close to those used in the previous work, but a better
agreement between experimental and analytical accelerations at the top of the model was
achieved if 3.5% of critical damping for the first and third modes was used. Therefore slightly
different damping parameters than those obtained from impact hammer test measurement were
adopted.
It can be concluded from the above observations that the FE model could capture reasonably
well the behaviour of the model for selected no-slip tests. A little difference between
experimental and FE model was found in the behaviour of the foundation interface, but this
difference was not significant and was explained.

191

APPENDIX A - PLOTS OF MODEL MOTIONS FROM NO-SLIP TESTS AND SIMULATIONS

a) Model Top Acceleration Relative to the Shake Table

Tirne^sec)
b) Model Base Acceleration Relative to the Shake Table

5.00
Q^

5.25

c) Shake Table Acceleration

5 50
Time (sec)

TO

CO
CD
O
O
<

5.00
QQ
4

5.25

d) Shake Table Displacement

5.50
Time (sec)

0.20

5.50
Time (sec)
Figure A . 1: Response of Dam Model to Harmonic Excitation at Frequency of 5 Hz
192

APPENDIX A - PLOTS OF MODEL MOTIONS FROM NO-SLIP TESTS AND SIMULATIONS

a) Model Top Acceleration Relative to the Shake Table

0.2

5.00

5.25

_ 5.50

5.75

Time (sec)

. b) Model Base Acceleration Relative to the Shake Table


0.4 i
n

5.00
Q

5.25

c) Shake Table Acceleration

5.50
Time (sec)

CD

E 0.00
CD

1-0.10
5.50
Time (sec)
Figure A.2: Response of Dam Model to Harmonic Excitation at Frequency of 7.5 Hz
193

APPENDIX

A - PLOTS OF MODEL MOTIONS FROM NO-SLIP

TESTS AND

SIMULATIONS

a) Model Top Acceleration Relative to the Shake Table

c
o

2
jD
CD
O
O
<

Time\sec)
_ .
0.4

b) Model Base Acceleration Relative to the Shake Table


- r

o
ro

0.0

JD
CD
O
O
<

-0.2

Experiment
FE Model

-0.4

0 ,|Q

5.00

5.25

5 50
Time (sec)

5.00

5.25

5.50
Time (sec)

d) Shake Table Displacement

5.75

6.0

0.05

5.50
Time (sec)

Figure A.3: Response of Dam Model to Harmonic Excitation at Frequency of 10 Hz


194

APPENDIX A - PLOTS OF MODEL MOTIONS FROM NO-SLIP TESTS AND SIMULATIONS

04

a) Model Top Acceleration Relative to the Shake Table

-0.2

5.00

5.25

_ 5.50
Time (sec)
b) Model Base Acceleration Relative to the Shake Table
x

_ .
0.4 -r

Q)

-0.2 -I

Experiment
FE Model

-0.4
5.00
0
3
c
o

c) Shake Table Acceleration

(0

-0.2-

<

5.75

6.0

5.75

6.0

0.0-

Ld)
_

5.50
Time (sec)

0.2 -

ccel

5.25

-04
5.00

5.25

d) Shake Table Displacement

5.00

5.25

5.50
Time (sec)

5.50
Time (sec)

Figure A.4: Response of Dam Model to Harmonic Excitation at Frequency of 12.5 Hz


195

APPENDIX A - PLOTS OF MODEL MOTIONS FROM NO-SLIP TESTS AND SIMULATIONS

a) Model Top Acceleration Relative to the Shake Table

_ .
0.4

_ 5.50 .
Time (sec)
b) Model Base Acceleration Relative to the Shake Table
-r

0.2
2

0.0

0)

-vwwwwwvw^
Experiment

3 -0.2

FE Model

-0.4
5.00
0

5.25

c) Shake Table Acceleration

5 50
Time (sec)

5.75

6.0

5.50
Time (sec)
Figure A.5: Response of Dam Model to Harmonic Excitation at Frequency of 15 H z
196

APPENDIX A - PLOTS OF MODEL MOTIONS FROM NO-SUP

TESTS AND SIMULATIONS

a) Model Top Acceleration Relative to the Shake Table

0.2

5.00

5.25

_ 5.50 .
Time (sec)
b) Model Base Acceleration Relative to the Shake Table

_ .
0.4 -r-

0.2
0.0
CD
O

3 -0.2

Experiment
FE Model

-0.4
5.00

5.25

_. 5.50 .
Time (sec)

5.25

5.50
Time (sec)

5.75

6.0

0.4 -,

"55
O
4*

0.20.0-

ccel

CD

<

-0.2 -04
5.00
)0

d) Shake Table Displacement

5.50
Time (sec)

Figure A.6: Response of Dam Model to Harmonic Excitation at Frequency of 17.5 Hz


197

APPENDIX A - PLOTS OF MODEL MOTIONS FROM NO-SUP TESTS AND SIMULATIONS

a) Model Top Acceleration Relative to the Shake Table

5.50
Time (sec)
_ . b) Model Base Acceleration Relative to the Shake Table
0.4 -T^
^

0.2
0.0

/VW\MA/VWWVWW\M

-0.2

Experiment
FE Model

-0.4
5.00
0

5.25

c) Shake Table Acceleration

Time (sec)

5.75

6.0

5.50
Time (sec)

Figure A.7: Response of Dam Model to Harmonic Excitation at Frequency of 20 Hz


198

APPENDIX A - PLOTS OF MODEL MOTIONS FROM NO-SLIP TESTS AND SIMULATIONS

a) Model Top Acceleration Relative to the Shake Table

ai 0.5

f.

'

*.

-1.0
5.00

5.25

5 50

5.75

6.0

5.50

5.75

6.0

5.50

5.75

6.0

Time (sec)

b) Model Base Acceleration Relative to the Shake Table

c) Shake Table Acceleration

Time (sec)

0.5

0.0

-I

1 -0.5-1
<
-1.0

5.00

5.25

d) Shake Table Displacement

5.00

5.25

Time (sec)

Time (sec)

Figure A.8: Response of Dam Model to Harmonic Excitation at Frequency of 22.5 Hz


199

APPENDIX A - PLOTS OF MODEL MOTIONS FROM NO-SUP

TESTS AND SIMULATIONS

a) Model Top Acceleration Relative to the Shake Table

\ r.K r.k

:.k

0.5
2

0.0

-0.5

-1.0
5.00

5.25

Time (sec)

5.75

6.0

b) Model Base Acceleration Relative to the Shake Table

1.0
^

0.5

0.0 4 /
-0.5 A

Experiment
FE Model

-1.0
5.00
1

5.25

c) Shake Table Acceleration

5.50
Time(sec)

5.75

6.0

0.5

5.00
QQ

5.25

d) Shake Table Displacement

5.50
Time (sec)

5.50
Time (sec)

Figure A.9: Response of Dam Model to Harmonic Excitation at Frequency of 25 Hz


200

APPENDIX B - SAMPLES OF INPUT FILES


FOR ANSYS ANALYSIS
A large number of numerical simulations using A N S Y S were performed during this study. In
order to optimize computing time the simulations were run in series using batch files. One
simulation typically consisted of preprocessing, solution and postprocessing phases, all being
controlled by commands from separate

files.

This appendix shows a sample of the

preprocessing file containing description of the studied system. In addition, a sample of the
solution file is shown, which contains A N S Y S commands to control the nonlinear time-domain
dynamics analysis.

S A M P L E OF A PREPROCESSING FILE:
!-- s e t f i l t e r s
KEYW,PR_SET,1
KEYW,PR_STRUC,1
KEYW,PR_THERM,0
KEYW,PR_ELMAG, 0
KEYW,PR_FLUID,0
KEYW,PR_MULTI, 0
KEYW,PR_CFD,0
KEYW,LSDYNA,0
/PMETH,OFF
i
preprocessor
/PREP7
!
D e f i n e PLANE STRESS element
ET,1,PLANE42!-- p l a n e q u a d r i l a t e r a l l i n e a r element
KEYOPT,1,1,0
KEYOPT,1,2,0
KEYOPT,1,3,3!-- K03=3 p l a n e s t r e s s w i t h t h i c k n e s s
KEYOPT,1,5,0
KEYOPT,1,6,0
i
r e a l c o n s t a n t s f o r p l a n e element
R,l,l.,
!-- t h i c k n e s s i n [m]
!__
m a t e r i a l p r o p e r t i e s f o r CONCRETE
M P , E X , 1 , 2 7 e 9 ! e l a s t i c modulus i n [Pa]
MP,EY,1,27e9!-- e l a s t i c modulus i n [Pa]
MP,DENS,1,2580!-- d e n s i t y i n [kg/m3]
MP,NUXY,1,0.22!-- P o i s s o n r a t i o
i
m a t e r i a l p r o p e r t i e s f o r FOUNDATION ROCK
MP,EX,2,15e9!-- e l a s t i c modulus
MP,EY,2,15e9
MP,DENS,2,10.!-- d e n s i t y i n [kg/m3] - almost massless f o u n d a t i o n
MP,NUXY,2,0.25!-- P o i s s o n r a t i o
i
D e f i n e CONTACT element
ET,2,CONTAC48
!-- node t o s u r f a c e c o n t a c t element
KEYOPT,2,1,0
KEYOPT,2,2,0
!--K03=2 r i g i d coulomb f r i c t i o n !CAUTION: g e t s u n s t a b l e ! s m a l l p i v o t .
KEYOPT,2,3,1!-- K03=l e l a s t i c coulomb f r i c t i o n
KEYOPT,2,7,1!-- K07=l use 'reasonable' time-step @ c o n t a c t time
i
r e a l c o n s t a n t s f o r c o n t a c t element
R,2,1.8e9,4ell, ,1.05, , , !-- KN, KT, ,FACT ( s t a t i c / k i n e t i c )

201

APPENDIX B - SAMPLES OF INPUT FILES FOR ANSYS ANALYSIS

i
m a t e r i a l p r o p e r t i e s f o r c o n t a c t element
MP,MU,3,1.!-- k i n e t i c f r i c t i o n c o e f f .
i
D e f i n e c o n c e n t r a t e d mass element
ET,3,MASS21
KEYOPT,3,3,0
! 6 components o f mass - 3 masses, 3 r o t a t o r y i n e r t i a
KEYOPT,3,2,0
! element coord system p a r a l e l t o g l o b a l coord system
added mass - Load Case 1, NHydro 1
R, 3, 201000.,0,0,0,0,0
added mass - Load Case 1, NHydro 2
R,4,282000.,0,0,0,0,0
added mass - Load Case 1, NHydro 3
R,5,176000.,0,0,0,0,0
added mass - Load Case 1, NHydro 4
R, 6,190000.,0,0,0,0,0
added mass - Load Case 1, NHydro 5
R, 7,173000.,0,0,0,0,0
added mass
R,8, 47000.,0,0,0,0,0
Load Case 1, NHydro 6
i
b u i l d model down up
key p o i n t s
9
10
K, 10,6.3000,45.000,0!-I top \ 8
K, 9, 0.000,45.000,0!-K, 8,13.560,34.000,0!-tx
K, 7, 0.000,34.000,0!-K, 6,20.820,23.000,0!-cg
\ DownStream=D
K, 5, 0.000,23.000,0!-UpStream=U
\
K, 4,28.740,11.000,0!-4
K, 3, 0.000,11.000,0!-K, 2,36.000, 0.000,0!-tx
K, 1, 0.000, 0.000,0!-x=0i
|y=0
i
f o u n d a t i o n rock
K,11,-20.000,
0.000,0!-K,12,-20.000,-20.000,0!-11--18----17
14
K,13, 56.000,-20.000,0!-|
f o u n d a t i o n rock
0.000,0!-12--15
16-K,14, 56.000,
-13
K,15, -3.111,-20.000,0
K,16, 39.111,-20.000,0
0.000,0
K,17, 39.111,
0.000,0
K,18, -3.111,
i
k e y p o i n t s on US f a c e f o r added masses
K,19, 0.000, 17.000,0!
f o r added mass Nhydro3
K,20, 0.000, 39.500,0!
f o r added mass Nhydro6
i
create h o r i z o n t a l l i n e s
L,l,2
!-- c r e a t e l i n e
CM,LineHl,LINE!-- c r e a t e a named component out o f l i n e
LSEL,NONE
!-- c l e a r s e l e c t i o n
L,3,4
CM,LineH2,LINE
LSEL,NONE
L,5,6
CM,LineH3,LINE
LSEL,NONE
L,7,8
CM,LineH4,LINE
LSEL,NONE
L,9,10
CM,LineH5,LINE
LSEL,NONE
L,ll,18
!-- base l e f t t o p
CM,LineFTl,LINE
LSEL,NONE
L, 18,17
! base mid t o p
CM,LineFT2,LINE
LSEL,NONE
L, 17,14
! base r i g h t t o p
CM,LineFT3,LINE
LSEL,NONE
L, 12,15
! base l e f t bottom
CM,LineFBl,LINE
LSEL,NONE

202

APPENDIX B - SAMPLES OF INPUT FILES FOR ANSYS ANALYSIS

L, 15,16
! base mid bottom
CM,LineFB2,LINE
LSEL,NONE
L, 16,13
! base r i g h t bottom
CM,LineFB3,LINE
LSEL,NONE
;
US f a c e v e r t i c a l l i n e s
L,l,3
CM,LineUl,LINE
LSEL,NONE
L,3,5
CM,LineU2,LINE
LSEL,NONE
L,5,7
CM,LineU3,LINE
LSEL,NONE
L,7,9
CM,LineU4,LINE
LSEL,NONE
L,12,11
CM,LineFVl,LINE! base l e f t v e r t i c a l
LSEL,NONE
L,15,18
CM,LineFV2,LINE! base m i d - l e f t v e r t i c a l
LSEL,NONE
;
DS f a c e v e r t i c a l l i n e s
L,2,4
CM,LineDl,LINE
LSEL,NONE
L,4,6
CM,LineD2,LINE
LSEL,NONE
L,6,8
CM, LineD3 , LINE
LSEL,NONE
L,8,10
CM,LineD4,LINE
LSEL,NONE
L,16,17
CM, LineFV3,LINE! base m i d - r i g h t v e r t i c a l
LSEL,NONE
L,13,14
CM,LineFV4,LINE! base r i g h t v e r t i c a l
LSEL,NONE
i
create
areas
KSEL,ALL
A, 1,2,4,3!-- bottom p a r t - t r a n s i t i o n
CM,Areal,AREA
ASEL,NONE
A,3,4,6,5
CM,Area2,AREA
ASEL,NONE
A, 5,6,8,7!-- second t r a n s i t i o n
CM,Area3,AREA
AS EL,NONE
A,7,8,10,9!-- t o p
CM,Area4,AREA
A, 12,15,18,11!-- bottom p l a t e - l e f t p a r t
CM,AreaFl,AREA
AS EL,NONE
A, 15,16,17,18!-- bottom p l a t e - mid p a r t
CM,AreaF2,AREA
AS EL,NONE
A, 16,13,14,17!-- bottom p l a t e - r i g h t p a r t
CM,AreaF3,AREA

203

APPENDIX B - SAMPLES OF INPUT FILES FOR ANSYS


ASEL,NONE

specify subdivisions
CMSEL,ALL
!- s e l e c t a l l components
LESIZE,LineHl,
6!-- s p e c i f y d i v i s i o n h o r i z o n t a l l i n e s
LESIZE,LineH2,
4
LESIZE,LineH3,
4
LESIZE,LineH4,
2
LESIZE,LineH5,
2
LESIZE,LineFTl
,2
LESIZE,LineFT2
,7
LESIZE,LineFT3
,2
LESIZE,LineFBl
,2
LESIZE,LineFB2
,7
LESIZE,LineFB3
,2
LESIZE,LineUl,
1!-- s p e c i f y d i v i s i o n a l o n g US f a c e
LESIZE, LineU2,
2
LESIZE,LineU3,
1
LESIZE,LineU4,
2
LESIZE,LineDl,
1!-- s p e c i f y d i v i s i o n a l o n g DS f a c e
LESIZE,LineD2,
2
LESIZE,LineD3,
l
LESIZE,LineD4,
2
i
mesh m o n o l i t h
element type p l a n e
TYPE,1
m a t e r i a l type 1 (concrete)
MAT, 1
r e a l contants 1
REAL,1
ESHAPE,2
!-- shape f o r c i n g t o quads i n u n i f o r m areas
AMESH,Area2
!-- mesh areas w i t h c o n c r e t e
AMESH,Area4
ESHAPE,3
-- mesh f o r t r a n s i t i o n areas
use quads i f p o s s i b l e .
AMESH,Areal
!-- mesh t r a n s i t i o n areas
AMESH,Area3
i
mesh rock f o u n d a t i o n
- element type 1
TYPE,1
MAT, 2
- m a t e r i a l type 2 ( f o u n d a t i o n rock)
REAL,1
- r e a l c o n t a n t s 1
ESHAPE,2
!-- shape f o r c i n g t o quads i n u n i f o r m areas
AMESH,AreaFl
!-- mesh areas w i t h rock f o u n d a t i o n - l e f t p a r t
AMESH,AreaF2
!-- mesh areas w i t h rock f o u n d a t i o n -mid p a r t
AMESH,AreaF3
!-- mesh areas w i t h rock f o u n d a t i o n - r i g h t p a r t
/PNUM,ELEM,1!-- d i s p l a y numbering
EPL0T!-- p l o t elements
generate elements w i t h added masses
TYPE,3 ! - element type MASS21
REAL,3!-- added mas - Load Case 1, NHydro 1 - k e y p o i n t 1
KMESH,1
TYPE,3 ! - element type MASS21
added mas - Load Case 1, NHydro 2 - k e y p o i n t 3
REAL,4 !KMESH,3
element type MASS21
TYPE,3 !added mas - Load Case 1, NHydro 3 - k e y p o i n t 19
REAL,5 !KMESH,19
TYPE,3 !element type MASS21
REAL,6 !added mas - Load Case 1, NHydro 4 - k e y p o i n t 5
KMESH,5
TYPE,3 ! - element type MASS21
REAL,7 !added mas - Load Case 1, NHydro 5 - k e y p o i n t 7
KMESH,7
TYPE,3 ! - element type MASS21
REAL,8 !added mas - Load Case 1, NHydro 6 - k e y p o i n t 20
KMESH,20
merge d u p l i c a t e nodes
NSEL,ALL
NUMMRG,NODES ,0.01!
t o l e r a n c e o f 0.01 m
NSEL,NONE

204

APPENDIX B-

SAMPLES OF INPUT FILES FOR ANSYS ANALYSIS

I
Setup c o n t a c t s
NSEL,S,NODE,,25,25
CM,Cont1,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,27,27
CM,Cont2,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,28,28
CM,Cont3,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,29,29
CM,Cont4,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,30,30
CM,Cont5,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,31,31
CM,Cont6,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,26,26
CM,Cont7,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
i
Setup t a r g e t s
NSEL,S,NODE,,43,43
NSEL,A,NODE,,62,62
CM,Targl,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,62,62
NSEL,A,NODE, ,63,63
CM,Targ2,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE, ,63,63
NSEL,A,NODE,,64,64
CM,Targ3,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,64,64
NSEL,A,NODE, ,65,65
CM,Targ4,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,65,65
NSEL,A,NODE,,66,66
CM,Targ5,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,66,66
NSEL,A,NODE, ,67,67
CM,Targ6,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
NSEL,S,NODE,,67,67
NSEL,A,NODE, ,59,59
CM,Targ7,NODE!-- make s e l e c t e d i n t o a component
NSEL,NONE
i
s e t props f o r c o n t a c t element and g e n e r a t e elem's
TYPE,2
MAT, 3
REAL,2
GCGEN,Contl,Targl!-- generate c o n t a c t elements
GCGEN,Cont2,Targ2
GCGEN,Cont3,Targ3
GCGEN,Cont4,Targ4
GCGEN,Cont5,Targ5
GCGEN,Cont6,Targ6
GCGEN,Cont7,Targ7
i
generate l a b e l s f o r nodes o f i n t e r e s t
NSEL,NONE
!
p o i n t s o f a p p l i c a t i o n o f hydro l o a d - numbered from bottom

205

APPENDIX B - SAMPLES OF INPUT FILES FOR ANSYS ANALYSIS

!NSEL,S,LOC,X, 0.000
!NSEL,R,LOC,Y, 0.000
! a t the bottom, k e y p o i n t 1
!*GET,NHydrol,NODE,,NUM,MIN!-- a s s i g n i t s number t o v a r i a b l e
!NSEL,NONE
NSEL,S,LOC,X, 0.000
NSEL,R,LOC,Y, 11.000
! a t 11 m from bottom, k e y p o i n t 3
*GET,NHydro2,NODE,,NUM,MIN!-- a s s i g n i t s number t o v a r i a b l e
NSEL,NONE
NSEL,S,LOC,X, 0.000
NSEL,R,LOC,Y, 17.000
! a t 17 m from bottom, k e y p o i n t 19
*GET,NHydro3,NODE,,NUM,MIN!-- a s s i g n i t s number t o v a r i a b l e
NSEL,NONE
NSEL,S,LOC,X, 0.000
NSEL,R,LOC,Y, 23.000
! a t 23 m from bottom, k e y p o i n t 5
*GET,NHydro4,NODE,,NUM,MIN!-- a s s i g n i t s number t o v a r i a b l e
NSEL,NONE
NSEL,S,LOC,X, 0.000
NSEL,R,LOC,Y, 34.000
! a t 34 m from bottom, k e y p o i n t 7
*GET,NHydro5,NODE,,NUM,MIN!-- a s s i g n i t s number t o v a r i a b l e
NSEL,NONE
NSEL,S,LOC,X, 0.000
NSEL,R,LOC,Y, 39.500
! a t 39.5 m from bottom, k e y p o i n t 20
*GET,NHydro6,NODE,,NUM,MIN!-- a s s i g n i t s number t o v a r i a b l e
NSEL,NONE
NSEL,S,LOC,X,-3.Ill !
t o p o f f o u n d a t i o n a t h e e l o f dam
NSEL,R,LOC,Y, 0.000
*GET,NStUs,NODE,,NUM,MIN
NSEL,NONE
NSEL,S,LOC,X,0.000 !-- t o p US c o r n e r o f m o n o l i t h
NSEL,R,LOC,Y,45.000
*GET,NToUs,NODE,,NUM,MIN
NSEL,NONE
NSEL,S,LOC,X,0.000
!-- bottom US c o r n e r o f m o n o l i t h i s a l s o NHydrol node
NSEL,R,LOC,Y,0.000
*GET,NBaUs,NODE,,NUM,MIN
NSEL,NONE
NSEL,S,LOC,X,36.000 !-- bottom DS c o r n e r o f m o n o l i t h
NSEL,R,LOC,Y,0.000
*GET,NBaDs,NODE,,NUM,MIN
NSEL,NONE
ALLSEL,ALL,ALL!-- s e l e c t e v e r y t h i n g
i
a p p l y boundary c o n d i t i o n s
KBC,1!-- stepped b o u n d a r y - c o n d i t i o n s a p p l i c a t i o n
CMSEL,S,LineFBl!-- s e l e c t l i n e s a l o n g u n d e r s i d e o f bottom p l a t e
CMSEL,A,LineFB2!-- add l i n e t o s e l e c t i o n
CMSEL,A,LineFB3!-- add l i n e t o s e l e c t i o n .
CMSEL,A,LineFVl!-- add l i n e t o s e l e c t i o n
CMSEL,A,LineFV4!-- add l i n e t o s e l e c t i o n
NSLL,S,1
!-- s e l e c t nodes a l o n g s e l e c t e d l i n e s
CM,CNSt,NODE
!-- make them i n t o a component
D,CNSt,UY,0
!-- r e s t r a i n t r a n s l a t i o n s i n y d i r e c t i o n .
D,CNSt,UX,0
!-- r e s t r a i n t r a n s l a t i o n s i n x d i r e c t i o n .
ALLSEL,ALL,ALL
EPLOT
i
a p p l y damping
ALPHAD,11.5 !-- v a l u e o f a l p h a f o r 1 s t and 3 r d modes and 2% o f damping
BETAD,0.000065 !-- v a l u e o f b e t a f o r 1 s t and 3 r d modes and 2% o f damping
ALLSEL,ALL,ALL!-- s e l e c t e v e r y t h i n g

206

APPENDIX B - SAMPLES OF INPUT FILES FOR ANSYS ANALYSIS

S A M P L E OF A SOLUTION FILE:
! s o l u t i o n f i l e t o s i m u l a t e response o f a 45 m t a l l f u l l s c a l e dam m o n o l i t h
i
define general s o l u t i o n c r i t e r i a
ANTYPE,4 !-- a n a l y s i s type: t r a n s i e n t dynamic
TRNOPT,FULL !-- f u l l s o l u t i o n
LUMPM,1 !-- lumped masses
NLGEOM,1 !-- l a r g e deformations
SSTIF,0 !-- no s t r e s s s t i f f e n n i n g
NROPT,AUTO!-- Newton-Rapson n o n l i n e a r i t i e s - auto s e l e c t i o n on
EQSLV,FRONT,1E-008,1 !-- s o l v e r s e l e c t i o n - f r o n t a l
CNVTOL,F,1.,.01,2
! s p e c i f y convergence c r i t e r i a
i
t r a n s i e n t a n a l y s i s i n t e g r a t i o n parameters
TINTP,0.005!-- " A r t i f i c i a l l y Damped" I m p l i c i t Newmark ( d e f a u l t )
i
l o a d s t e p 1: g r a v i t y
TIMINT,0!-- i g n o r e i n e r t i a e f f e c t s
TIME,0.005!-- s e t end time f o r g r a v i t y l o a d a p p l i c a t i o n
AUT0TS,1!-- automatic time s t e p o f f
DELTIM,0.001,0.0001,0.005,ON!-i n i t i a l time s t e p
ACEL,0,9.807,0,!-- a p p l y g r a v i t y a c c e l e r a t i o n l o a d
OUTRES,NSOL,LAST
!set output frequency
/SOLU!-- s w i t c h t o s o l v e r
SOLVE!-- r u n s o l v e r
i
l o a d step 2: hydro
TIMINT,0!-- i g n o r e i n e r t i a e f f e c t s
TIME,0.01!-- s e t end time
AUTOTS,l!-- automatic time s t e p o f f
DELTIM,0.001,0.0001,0.005,ON!-i n i t i a l time step
!F,NHydrol,FX,2180000.!-- a p p l y hydro l o a d NHydrol (bottom node)
F,NBaUs,FX,2180000.!-- a p p l y hydro l o a d NHydrol (bottom node=NBaUs)
F,NHydro2,FX,2790000.!-- a p p l y hydro l o a d NHydro2
F,NHydro3,FX,1540000.!-- a p p l y hydro l o a d NHydro3
F,NHydro4,FX,1430000.!-- a p p l y hydro l o a d NHydro4
F,NHydro5,FX,1000000.!-- a p p l y hydro l o a d NHydro5
F,NHydro6,FX, 204000.!-- a p p l y hydro l o a d NHydro6
OUTRES,NSOL,LAST!-- output r e s o l u t i o n
/SOLU!-- s w i t c h t o s o l v e r
SOLVE!-- r u n s o l v e r
i
l o a d s t e p s 3,4,5,....: base a c c e l e r a t i o n
TIMINT,1!-- i n c l u d e i n e r t i a e f f e c t s
KBC,0!-- use ramped BC a p p l i c a t i o n
Tm_Sta=0.015!-- s t a r t time (= end time o f f i r s t l o a d step)
Tm_End=2 0.!-- end time - FULL
Tm_lnc=0.005!-- l o a d s t e p d u r a t i o n
AUT0TS,1!-- automatic time s t e p on
!-1,000Hz 20,000Hz
200Hz
s o l u t i o n frequency l i m i t s
DELTIM,0.001,0.00005,0.005,ON!-time step ( t r y t o keep 5 t s / l s )
[ s p e c i f y : i n i t i a l , min, max, c a r r y from p r e v i o u s LS
OUTRES,NSOL,LAST!-- s e t output frequency t o l a s t substep o f each
!
l o a d s t e p f o r a t o t a l o f 4010 r e c o r d s
!
read i n a c c e l e r a t i o n s
*DIM,Accel,TABLE,4010,1!-- d e f i n e a c c e l e r a t i o n t a b l e
*VREAD, A c c e l (0,0) , time200, p m , , 1! - - read time from f i l e
(F8.4)!-- format
*VREAD,Accel(0,1),eqlt200,prn,,1!-- read a c c e l e r a t i o n s from f i l e
(F8.4)!-- format
i
a p p l y and s o l v e f o r each l o a d s t e p
*DO,Tm,Tm_Sta,Tm_End,Tm_Inc!-- f o r a l l l o a d s t e p s
TIME,Tm!-- s e t time
AccSca=Accel(Tm,l)*9.807*0.453
ACEL,AccSca,9.807,0!-- a p p l y base a c c e l e r a t i o n
SOLVE!-- s o l v e
*ENDDO!-- next l o a d s t e p
SAVE

207

APPENDIX C - PLOTS FROM F E AND S D O F


SIMULATIONS ON A FULL-SCALE D A M M O N O L I T H
Two numerical models were used in Chapter 6 to simulate response of a full-scale dam monolith
due to three different types of earthquake motions. The dam monolith unbonded at the base was
subjected to the base excitations of varying PGA. The water level of the reservoir was another
parameter varied during the excitations and a total of nine water levels from 60% to 100% of the
full reservoir were considered. The amounts of sliding of the monolith obtained from SDOF
analyses were compared with those using the F E model. Another way to test the numerical
models is to compare the experimental results with those from the simulations in time domain
and it is done in this appendix.
A total of six simulations were selected for analysis in this appendix. They include three with
PGA of 1.04g and three with PGA of 0.79g. These PGA's were used with all earthquake records
considered in Chapter 6. The simulated water level was 100% full reservoir for all simulations.
The comparison of results is performed in Figures D . l to D6. Each of these figures contains
four parts a to d with plots of the obtained results. Parts a present time histories of dam monolith
relative displacement (sliding) with respect to the base. It can be observed from these plots that
the time histories from SDOF and F E analyses follow very similar trends. However, it cannot
be concluded that one of the models would consistently produce more or less sliding than the
other.
Parts b and c of all figures contain plots of the relative accelerations of the dam monolith with
respect to the base. The FE-based acceleration was obtained at the point on the upstream side
of the base of the monolith. Only one acceleration record is obtained from the SDOF
simulations and it is used for the comparison in this appendix. The character of the acceleration
time histories has some similar as well as some different features for the F E and SDOF-based
records. The peak relative accelerations occur about at the same time for all the records, but the
FE-based peak accelerations are higher in most cases. In addition, the F E accelerations have
non-zero values throughout the records. This is in contrast with the SDOF-based accelerations,
which show zero relative accelerations except of the point where the dam slid during the
analysis. These observations are in agreement with expected behaviour of the models. The F E
model yielded non-zero accelerations throughout the records because of the flexibility in
horizontal direction of the contact elements used to simulate the dam to base contact. No such
flexible elements were used in the SDOF model.
The plots in this appendix indicate that foundation interfaces of both numerical models used in
the study do not perform the same way. However, it cannot be concluded that this would affect
the amount of sliding of the model in such a way that one model would be consistently
conservative or unconservative when compared with the other. Therefore, it can be concluded
that both models can be used for the analysis if the objective of it is to estimate the amount of
sliding of the dam monolith with respect to its base.
208

APPENDIX C - PLOTS FROM FE AND SDOF SIMULATIONS ON A FULL-SCALE

DAM MONOLITH

a) Dam Downstream Sliding - SDOF and FE Models

6
4
FE Model

SDOF Model

0
1

Time (sec)

r
6

b) Dam Base, Relative Acceleration - FE Model

1 2

Time (sec)

c) Dam Base Relative Acceleration - SDOF Model

Time (sec)

d) Ground Acceleration

'

.6

Time (sec)

Figure C . l : Response of Dam Monolith to Nahanni Earthquake Record with P G A 1.04g


209

APPENDIX C - PLOTS FROM FE AND SDOF SIMULATIONS ON A FULL-SCALE

DAM MONOLITH

a) Dam Downstream Sliding - SDOF and FE Models

FE Model
SDOF Model

4
Time (sec)

4
2

b) Dam Base Relative Acceleration - FE Model

12

Time (sec)

c) Dam Base Relative Acceleration - SDOF Model

Time (sec)

d) Ground Acceleration

4 _

, .6

Time (sec)

Figure C.2: Response of Dam Monolith to Nahanni Earthquake Record with P G A 0.79g

210

APPENDIX C - PLOTS FROM FE AND SDOF SIMULATIONS ON A FULL-SCALE

20

DAM MONOLITH

a) Dam Downstream Sliding - SDOF and FE Models

15
=
<D
10
E
o
TO 5

FE Model

D.
CO

SDOF Model

0
10~

Time^sec)

30

1.2 -,
0.6 c
o
2 0.0 CD
<D
-0.6 -

<

-1.2
C

Time (sec)

1.2 -.
3
0.6c
o
2 0.0
CD
CD
-0.6-

<

-1 ?
0

10

Time^sec)

. _ d) Ground Acceleration
i
-i

Time (sec)
Figure C.3: Response of Dam Monolith to Mexico Earthquake Record with P G A 1.04g
211

APPENDIX C - PLOTS FROM FE AND SDOF SIMULATIONS ON A FULL-SCALE

DAM MONOLITH

a) Dam Downstream Sliding - SDOF and FE Models

2Q

20

men

15
<D
Q

(0
CL
V)

10
FE Model

5
0 -

10

30

b) Dam Base Relative Acceleration - FE Model

1 2

-ft

Time (sec)

SDOF Model

0.6

Time (sec)

1.2
3

c
o
2

0.60.0

8 -0.6-

<

-1.2
C

Time (sec)
d) Ground Acceleration

Time^sec)

30

Figure C.4: Response of Dam Monolith to Mexico Earthquake Record with P G A 0.79g
212

APPENDIX C - PLOTS FROM FE AND SDOF SIMULATIONS ON A FULL-SCALE

DAM MONOLITH

a) Dam Downstream Sliding - SDOF and FE Models

15

E
c

E
cu
o
ro
o_
w

10
FE Model

SDOF Model
0
Time (sec)
b) Dam Base Relative Acceleration - FE Model

0.6

0.0

8
<

-0.6

0)
<D

I"

-1.2
2

c
o

4I

I
Time^sec)

d) Ground Acceleration

1.2

Time%ec)

c) Dam Base Relative Acceleration - SDOF Model

1.2

c
o

0.6

2 0.0
03
0)
O
O -0.6

<

-|-. 6.
.
Time (sec)

Figure C.5: Response of Dam Monolith to Northridge Earthquake Record with P G A 1.04g
213

APPENDIX C - PLOTS FROM FE AND SDOF SIMULATIONS ON A FULL-SCALE

DAM MONOLITH

a) Dam Downstream Sliding - SDOF and FE Models

10
5

FE Model
SDOF Model

0
Time (sec)

1.2 -, b) Dam Base Relative Acceleration - FE Model


0.6 c
o
E 0.0aC>
uoD -0.6-

<

-1 ? Time (sec)
c) Dam Base Relative Acceleration - SDOF Model

1.2
0.6
E

0.0

jD
CD

8 -0.6
-1.2
2
1.2 -,
c
o

ro
i

CD
CD
O
O

<

n
4i

Time i(sec)

d) Ground Acceleration

0.6 -

0.0-0.6-1 ?

Time (sec)

Figure C.6: Response of Dam Monolith to Northridge Earthquake Record with P G A 0.79g
214

APPENDIX D - CALCULATIONS FOR RELIABILITY


STUDY USING RESULTS OF F E SIMULATIONS
(Analysis using Mathcad 8 Professional - Mathsoft, 1998)
Characteristics o f Peak Ground Acceleration ( P G A ) distribution:
M e a n v a l u e o f P G A (in g ' s ) :
Coefficient of variation:

a M = n.l
v a G := 0.6

R e l a t i v e w a t e r l e v e l s i n % o f full r e s e r v o i r :
Import results from S D O F simulations:

j = 1.. 9
:

Indexes:

F E . . := F E 1 . .
i,k
i,k

Determine Parameters o f Linear Interpolation


(for e a c h w a t e r l e v e l ) :

FE. . . := F E 2 . ,
1 + 4,k
i,k
a

; =

n := 1.. 3

k := 1 . . 10 m := 1.. 12

FE3 := READPRN(filein)

FE. , := F E 3 . ,
1+8,k
i,k

intercept^FE<i>, FE"
^\>)

: = s l o p e

FE
x : = F E , , . F E , , +

"1.1'

1.1

absolute terms:

0
-1

4.907
4.767

a = 1.518

-2

b = 4.558

1.731

-3

4.440

1.961

-4

4.283

2.055
0.1

b)Sliding vs. PGA Plot - Water Level 6 5 %

Ln (PGA (g))

4.167

2.204

Ln (PGA (g))

-0.3

<J+'>

4.999

1.386

-0.5

) F E

5.195

1.241

-0.1

<i>

coefficients o f linear terms:

0.970

-0.3

F E

- FE,
..FE
'4,1
10

1.100

-0.5

n.l

a)Sliding vs. PGA Plot - Water Level 6 0 %


1

-0.7

: =

filein := c o n c a t ( v \ identif, "FElr.prn')

-5

F E 2 := READPRN^fileinJ

Linear interpolations o f results f r o m S D O F simulations


f o r e a c h w a t e r l e v e l , all e a r t h q u a k e s c o n s i d e r e d :

i = i.. 4

h. := 60 + ( j - 1 ) 5

identif := num2st<n)

FE1 := R E A D P R N ( f i l e i i j )
Put all r e s u l t s in o n e m a t r i x :

Arrival rate (quakes per year):

3.950

c)Sliding vs. PGA Plot - Water Level 7 0 %

-0.9

-0.7

-0.5

-0.3

Ln (PGA (g))

215

APPENDIX D - CALCULATIONS FOR RELIABILITY STUDY USING RESULTS OF FE SIMULATIONS

Sliding vs. PGA Plot - Water Level 75%

Sliding vs. PGA Plot - Water Level 80%


1
1
1
1

+_

- -

+-

-1-

~ +

"0.5

-0.3

0.9

"0.7

"0.5

-0.3

Ln (PGA (g))

-0.1

0.1

Ln (PGA (g))

Sliding vs. PGA Plot - Water Level 85%


1 ' 1
1

Sliding vs. PGA Plot - Water Level 90%

+_

+-

+
~r

_ +

*0.9

"0.7

-0.5

"0.3

-0.1

0.1

"0.5

Ln (PGA (g))

"0.3

Ln (PGA (g))

Sliding vs. PGA Plot - Water Level 95%


1
1
1
1
-t-

Sliding vs. PGA Plot - Water Level 100%


1
1
1

jfc^^

00

+
~

+
- +

0.9

-0.7

"0.5

"0.3

-0.1

Ln (PGA (g))

0.9

1
"0.7

"0.5

"0.3

-0.1

0.1

Ln (PGA (g))

Figure D1. Sliding vs. PGA Relationship for Different Water Levels Using Data from FE Analyses

216

APPENDIX D - CALCULATIONS FOR RELIABILITY STUDY USING RESULTS OF FE SIMULATIONS

Linear interpolations for all water levels, all earthquakes considered:


Linear interpolation of "a" coefficients:

Linear interpolation of "b" coefficients:

al := intercepts,a)

al = -0.982

bl := intercept,b)

bl = 7.000

a2 := slope(h,a)

a2 = 0.032

b2 := slope(h,b)

b2 =-0.030

Export coefficients: WRITER "FE") := al

50

60

70

80

90

APPENLX"FE") := a2

100

110

APPEND("FE") := bl

""50

60

70

Water level h (%)

80

APPEND("FE") := b2

90

100

110

Water level h (%)

Fig. D2: Linear Regression of 'a' coefficients

Fig. D3: Linear Regression of 'b' coefficients

Setup the interpolation function as follows: ln(D) = a(h)+b(h).ln(aG) or: ln(D) = a1+a2.h+(b1+b2.h).ln(aG)
lnD(lnaG,hh) := al + a2hhi- (bl + b2hh)lnaG

Calculate values of displacements (sliding) using this function:


FORMm, .1:= FEm, .1

F 0 R M

m,j,. =
:

l n D

F E

m,.- j,
h

Calculate matrix of errors: errorm,j. := FEm,j+-. 1. - FORMm,j+. 1,


Convert the matrix into a vector:
Put values of natural logarithms of PGA into a vector:

error - vei
9-(m-

i)+j

v e (

] Q
na

ve(

l)-l-m

:=

F E

error

m,l

Statistics for two random variables: natural logarithm of peak ground acceleration (InaG)
and the errors between the results from simulations and from interpolation formula:
Standard deviation of In(aG)

(distribution considered):

, _ J, (j

Mean value of In(aG):

i G_m:= ln(aM) - I ln(l + VaG

Mean error:

m:= mean

lnaG_a= 0.555

2
2

na

Standard deviation of errors:


E a=0.790

a:=

naG

error

InaG m=-2.456

_m = -1.336-10' .-17

N i st(error vec)
:=

2-e o= 1.580

a:

error vec - m

i=^l1

217

APPENDIX D - CALCULATIONS

FOR RELIABILITY STUDY USING RESULTS OF FE SIMULATIONS

Correlation coefficient between error_vec and distribution of InaG:


1

p:=
N

InaG vec - InaG m) (error vec - 8_m)]


e

InaG a-E <j

i= 1

Plot showing comparison


result from simulations with
those using interpolation
formula. The dashed lines
represent the range of
+/- 2 times standard
deviation of the errors.

-3

p = -9.969-10'

Q
E

-5

+t^

-1

-3

_L
3

Ln(D) from linear formula

Fig. D4: Comparison of Sliding from FE simulations with values using interpolation function

Plot showing distribution


of the errors against the
values of logarithms of
the peak ground
accelerations.

|
w

Fig. D5: Distribution of errors against values of PGA


Reliability analysis with 2 random variables and neglecting the lower bound of Ln(aG) distribution:
Setup performance function G: Mean of G:
Standard deviation of G:

G_m(S0,h)

G_c(h)

:=

ln(S0) -

(al + a2h + (bl + b2h)lnaG_m+ e_m)

:= ,J(bl + b2-h) lnaG_o + 2 ( b l + b2 h) p InaG jy E_CT + E C


2

218

APPENDIX D-

CALCULATIONS FOR RELIABILITY STUDY USING RESULTS OF FE SIMULATIONS

Allowable displacements (sliding) considered:


Reliability indexes:

S0:=(1 2 5 10 15 20)

G_m(S0.,h.)

Pj.d

4.896-10"

G c(h

Annual Probabilities of failure:


3.61710"

Pe. := cnorm(-p.

J/ Event probabilities of failure:

d:= 1.. last(S0)

1.327 10"
1.774 10'

j.d '

l-

exD^-v-Pe
\
j.d
F

3.253 10" 1.057 10" 5.347 10" 3.26110"

4.262 10"

1.360 10" 6.793 10" 4.104 10"

Water levels (%)


60%
65%

6.712-10" 2.404 10"

5.660 10" 1.772 10" 8.741 10" 5.231 10"

70%

9.326-10" 3.303 10"

7.623 10" 2.341 10" 1.14010" 6.758 10"

75%

1.042 10" 3.140 10" 1.510 10" 8.860 10"

1.448 10" 4.281 IO" 2.032 10"

80%
85%

Pa = 1.314 10" 4.606 10"


5

1.880 10" 6.522 10"


5

2.729-10" 9.387 IO"


5

4.025 10"

1.374 10"

[6.034 10" 2.048 IO"


5

1.180 10"

90%

2.047 10" 5.937 10" 2.781 10

A JI

J ^99 l Q~ ^
2.946 10" 8.388 10" 3.878 10" 2.207 10"

95%

4.322 10" 1.209 10" 5.51710" 3.108 10"

100%

i r\~

r\f> m

-t r\"

7 /

1 r\~

15

10
Sliding 1 cm: p S01.:= Pa<i>
a

j
p S05. := Pa<3>

Level of Sliding: 1
Divide matrix
of probabilities
into columns for each
level of sliding.

J.d

20 cm
Sliding 2 cm:

p S02.
a

Pa<2>

Sliding 5 cm:

Sliding 15 cm: p S15.:= Pa<5>


a

110

Sliding 10 cm: p si0. = Pa<4>


a

Sliding 20 cm: p S20. := Pa<6>


a

Annual Probabilities of Failure

PaSOl.

The annual probabilities


of failure represent the
probability of exceeding
the given permanent
displacement of the
dam. These displacement
levels were specified above.
The probabilities of failure
are shown on logarithmic
scale.

j 110

PaS02
J

PaS05
1 10
:

PaSlO.
J
PaS15j 1 -10
PaS20.

110

110

Water Level (%)

Fig. D6: Annual Probabilities of Exceeding a Given Amount of Sliding


219

APPENDIX E - CALCULATIONS FOR RELIABILITY


STUDY USING RESULTS OF S D O F SIMULATIONS
(Analysis using Mathcad 8 Professional - Mathsoft, 1998)
Characteristics of Peak Ground Acceleration (PGA) distribution:
Mean value of PGA (in g's):
Coefficient of variation:

M:=0.1

vaG := 0.6

Relative water levels in % of full reservoir:

Arrival rate (quakes per year):

j := 1.. 9

Import results from SDOF simulations: i d e n t i p num2st<j)


SDj := R E A D P R N [ f i l e i n _
10

endj := max^SDxj

j}

f i l e i p concat("ew", identif, "s.prn"

<2>

SDy = l i | [ SD

SDx^lnj^p]

start := mir^SDXj

0.1

hj := 60 V (j - 1)5

j:

lengths := rows( SDXj

. , . ,
, max(end) - mir( start) \
x := mir(start), (mir(start) +
-,
].. max(end)
Determine Parameters of Linear Interpolation

(for each water level):


level)

.
J

: =

j tercep( (SDx),
n

Linear interpolations of results from SDOF simulations


for each water level, all earthquakes considered:

-0.2
L n (PGA

0.2

i)\

b. :=. slope! fSDx.^, fSDy.^'

absolute terms: coefficients of linear terms

a)Sliding vs. P G A Plot - Water Level 60%

-0.4

fSDy.) 1

V'K

-0.674

6.628'

-0.491

6.949

-0.185

6.970

0.176

7.114

0.578

b = 6.734

1.030

6.479

1.546

5.989

1.665

5.887

2.060

5.648

0.4

(g))

b)Sliding vs. P G A Plot - Water Level 65%

c)Sliding
Plot - Water
Level 70%
p B vs. !P G A
!

~ L

C/5
C

-0.4

-0.2
L n (PGA

(g))

0.2

0.4

-0.4

-0.2
Ln (PGA

0.2

0.4

(g))

220

APPENDIX E - CALCULATIONS FOR RELIABILITY STUDY USING RESULTS OF SDOF SIMULATIONS

Ln (PGA (g))

Ln (PGA (g))

Ln (PGA (g))

Ln (PGA (g))

Figure E1. Sliding vs. PGA Relationship for Different Water Levels Using Data from SDOF Analyses

221

APPENDIX E - CALCULATIONS FOR RELIABILITY STUDY USING RESULTS OF SDOF SIMULATIONS

Linear interpolations for all water levels, all earthquakes considered:


Linear interpolation of "a" coefficients:

Linear interpolation of "b" coefficients:

al := intercept;h,a)

al =-5.158

bl := intercept, b)

bl =9.076

a2 := slope(h,a)

a2 = 0.072

b2 := slope(h,b)

b2 =-0.032

Export coefficients: WRITE("SD") := al

50

60

70

80

90

100

APPEND("SD") := a2

110

APPEND("SD") := bl APPEND("SD") := b2

50

60

Water level h (%)

70

80

90

100

110

Water level h (%)

Fig. E2: Linear Regression of 'a' coefficients

Fig. E3: Linear Regression of 'b' coefficients

Setup the interpolation function as follows: ln(D) = a(h)+b(h).ln(aG) or: ln(D) = a1+a2.h+(b1+b2.h).ln(aG)
lnD(lnaG,hh) := al + a2hh+ (bl + b2hh)lnaG
Calculate values of displacements (sliding) using this function:

Df := lnD^SDx.,h.)

Stack the values from simulations and those from interpolations into vectors:
lnaG_vec:= stack^SDx,, SDxjj

lnaG_vec = stack ^lnaG_vec, SDxjj

lnaG_vec:= stack ^lnaG_vec, SDx

lnaG_vec:= stack ^lnaG_vec, SDx^

lnaG_vec = stack (^lnaG_vec, SDx^

lnaG_vec:= stack ^lnaG_vec, SDx

lnaG_vec:= stack ^lnaG_vec, SDx^

lnaG_vec = stack ^lnaG_vec, SDx^

SDy_vec := stack (SDy,, SDy )

SDy_vec = stack (SDy_vec, SDy )

SDy_vec:= stack (SDy_vec, SDy

SDy_vec:= stack ^SDy_vec, SDy )

SDy_vec = stack (SDy_vec, SDy )

SDy_vec:= stack (SDy_vec, SDy

SDy_vec:= stack (SDy_vec, SDy )

SDyvec = stack (SDy_vec,SDy )

SDf_vec:= stack(SDf,, SDfj)

SDf_vec: = stack(SDf_vec, SDij)

SDf_vec:= stack (SDf_vec, SDf )

SDf_vec:= stack (SDfvecSDfj)

SDf_vec: = stack(SDf_vec, SDfg)

SDf_vec:= stack (SDf_vec, SDf,)

SDf_vec:= stack (SDf.vec.SDfg)

SDf_vec: = stack(SDf_vec,SDf )

N := rows(lnaG_vec)

n = 1..N

222

APPENDIX E - CALCULATIONS

FOR RELIABILITY STUDY USING RESULTS OF SDOF SIMULATIONS

Statistics for two random variables: natural logarithm of peak ground acceleration (InaG)
and the errors between the results from simulations and from interpolation formula:
Standard deviation of In(aG)
^ ln(l1 i--h'VaG'
InaG a= 0 . 5 5 5
InaG_rj:=
a:= Jin
(distribution considered):
Mean value of In(aG):
Calculate vector of errors:

lnaG_m:= ln(aM) - l - l n ( l + VaG )


2
error_vec := SDf_vec- SDy_vec

Mean error:

lnaG_m= - 2 . 4 5 6

m : =

-3

mean(error_vec)

Standard deviation of errors:

E m = -3.21210"

N
a:

E a =0.791

error_veCj - _m

2 E <J = 1.582

i= 1

Correlation coefficient between error_vec and distribution of InaG:


N

1
1

InaG vec - InaG

Plot showing comparison


result from simulations with
those using interpolation
formula. The dashed lines
represent the range of
+/- 2 times standard
deviation of the errors.

m)

(error vec

p = 4 . 8 3 5 10"

InaG O E a

- E m

Ln(D) from linear formula

Fig. E4: Comparison of Sliding from SDOF simulations with values using interpolation function
4r

Plot showing distribution


of the errors against the
values of logarithms of
the peak ground
accelerations.
-0.8

-0.6

-0.4

-0.2

0.2

0.4

Ln(aG) values

Fig. E5: Distribution of errors against values of PGA


Reliability analysis with 2 random variables and neglecting the lower bound of Ln(aG) distribution:
Setup performance function G: MeanofG:
Standard deviation of G:

G_m(S0,h) := l n ( S 0 ) - (al -i-a2h + (bl + b2-h)-lnaG_m*

e_m)

G_rj(h) := J(bl + b 2 h ) InaGjj + 2-(bl + b2h)plnaG_o_ai- z_c


2

223

APPENDIX E - CALCULATIONS FOR RELIABILITY STUDY USING RESULTS OF SDOF SIMULATIONS

Allowable displacements (sliding) considered


Reliability indexes
R
p

j,d-

G_m(S0.,h.,
V " J/
G ofhA

S0:=(1 2 5 10 15 20)

Event probabilities of failure:

4.274-10"

p.
a

1.860 10" 5.919 10"

2.405 10"

1.400 10"

3.679 10" 2.134 10"

1.420 10"

1.655 10" 7.158 10" 2.244 10"

1.042 10" 4.51210"


6

2.661 IO"

1.15010"

4.327 10"

1.871 IO" 5.835 10'

5.148 10"

Divide matrix
of probabilities
into columns for each
level of sliding.

1.440-10"

75%

8.967 10" 5.162 10" 3.464 10"

1.430-10"

8.204 10" 5.491 10"

80%
85%

2.31410"

1.324 10" 8.838 10"

90%

3.800 10'

2.169 10"

1.445 10"

95%

6.338 10" 3.610 10" 2.401 10"

100%

P a S

20 cm

15

o i . = (p * <i>
:

p S05. := Pa
J
p S!5. := (pa<5>
a

<3>

Sliding 5 cm:

65%

9.482 10"

10

Sliding 1 cm:

60%

Water levels (%)


9

5.704 10" 3.295 10" 2.21810"

1.605 10"

Allowable Sliding: 1

70%

3.593 10"

7.121 IO" 3.084 10" 9.609 10"


5

:= 1 - expf-v-Pe.
J.d

= cnorW-B
'

' '

1.592-10" 9.311 10" 6.326 10"

1.18510"

1.216 10" 3.895 10"

6.634 10" 2.879-10" 9.109 10"

Pa:

p
j

Annual Probabilities of failure:


2.785-10'

d := 1.. last(S0)

Sliding 2 cm:

p S02 := Pa<2>
a

Sliding 10 cm: p si0

Pa

Sliding 20 cm:

Pa

<4>

Sliding 15 cm:

p s20
a

<6>

Annual Probabilities of Failure

110
PaSOl

The annual probabilities


of failure represent the
probability of exceeding
the given permanent
displacement of the
dam. These displacement
levels were specified above.
The probabilities of failure
are shown on logarithmic
scale.

PaS02

110

PaSOS
_Jl-10

PaSlOj
PaS15jM0
PaS20
110

110

110

Water Level (%)

Fig. E6: Annual Probabilities of Exceeding a Given Amount of Sliding


224

You might also like