Vous êtes sur la page 1sur 394

Copyright

by

MEHMET BARIS DARENDELI

2001
DEVELOPMENT OF A NEW FAMILY OF NORMALIZED

MODULUS REDUCTION AND MATERIAL DAMPING

CURVES

by

MEHMET BARIS DARENDELI, B.S., M.S.

DISSERTATION

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

DOCTOR OF PHILOSOPHY

The University of Texas at Austin

August, 2001
Dedicated

To

My Parents,

My Wife and My Daughter


Acknowledgements

I would like to thank my supervising professor Dr. Kenneth H. Stokoe, II

for his guidance and support through the course of this study. His passion and

enthusiasm in his work has always inspired me. Our stimulating conversations

have made this study enjoyable.

Dr. Robert B. Gilbert’s assistance and guidance, which have made this

dissertation possible, is gratefully acknowledged. Besides his valuable input to

this work, he has influenced my perception of science and engineering with his

lectures on decision, risk and reliability.

I would also like to thank my dissertation committee members Dr. Jose M.

Roesset, Dr. Ellen M. Rathje, Dr. Alan F. Rauch and Dr. Mark F. Hamilton for

reviewing this dissertation in such a limited time frame and for their valuable

contributions to this work. Thanks are also extended to the rest of the former and

current geotechnical engineering faculty, Dr. Roy E. Olson, Dr. David E. Daniel,

and Dr. Stephen G. Wright for their lectures that broadened my knowledge.

The support from the California Department of Transportation, the

National Science Foundation, the Electric Power Research Institute, and Pacific

Gas and Electric Company is gratefully acknowledged for funding various stages

of the ROSRINE project. I would also like to acknowledge the contributions of

the National Institute of Standards and Technology, the United States Geological

Survey, the Department of Energy, the Westinghouse Savannah River

Corporation, Kajima Corporation, Geovision, Agbabian Associates, Fugro, Inc.,

Earth Mechanics, Inc., S&ME, Inc. in funding the research projects the results of

v
which are utilized in this study. Encouragement and guidance from Dr. Clifford

Roblee, Dr. John Schneider, Dr. Walter Silva, Dr. Robert Pyke, Dr. Robert

Nigbor, Dr. David Boore, Prof. Mladen Vucetic and Dr. Richard Lee, who took

part in these research projects, are appreciated.

Thanks to my best friend Cem Akguner for always being there whenever I

needed him, to Dr. Brent L. Rosenblad for trying to teach me how to bat

whenever we overworked, to Dr. Ahmet Yakut for our stimulating card plays and

arguments regarding them that lasted for hours, and to Baris Binici for each and

every five minute coffee break at 100oF. You have kept me sane (although

everyone reading this paragraph will question it a little) for the past seven years.

I would also like to thank the former and current graduate students that I

have worked side by side. I enjoyed each and every day and night that I worked

together with Dr. James A. Bay, Dr. Seon-Keun Hwang, Farn-Yuh Menq, Brian

Moulin, Celestino Valle and Nicola Chiara. Thanks are also extended to other

graduate students of whom I had the pleasure of making acquaintance; Dr. Eric

Liedtke, Dr. Mike Kalinski, Jeffrey Lee, Paul Axtell, Jiun Chen, Cem Topkaya
and many others that I unfortunately omitted. I would also like to thank Teresa

Tice-Boggs and Alicia Zapata for their administrative support, and Frank Wise,

Gonzalo Zapata, Max Trevino and Paul Walters for their technical assistance over

the years.

vi
DEVELOPMENT OF A NEW FAMILY OF NORMALIZED

MODULUS REDUCTION AND MATERIAL DAMPING

CURVES

Publication No._____________

Mehmet Baris Darendeli, Ph.D.

The University of Texas at Austin, 2001

Supervisor: Kenneth H. Stokoe, II

As part of various research projects [including the SRS (Savannah River Site)

Project AA891070, EPRI (Electric Power Research Institute) Project 3302, and

ROSRINE (Resolution of Site Response Issues from the Northridge Earthquake)


Project], numerous geotechnical sites were drilled and sampled. Intact soil

samples over a depth range of several hundred meters were recovered from 20 of

these sites. These soil samples were tested in the laboratory at The University of

Texas at Austin (UTA) to characterize the materials dynamically. The presence of

a database accumulated from testing these intact specimens motivated a re-

evaluation of empirical curves employed in the state of practice. The weaknesses

of empirical curves reported in the literature were identified and the necessity of

vii
developing an improved set of empirical curves was recognized. This study

focused on developing the empirical framework that can be used to generate

normalized modulus reduction and material damping curves. This framework is

composed of simple equations, which incorporate the key parameters that control

nonlinear soil behavior. The data collected over the past decade at The University

of Texas at Austin are statistically analyzed using First-order, Second-moment

Bayesian Method (FSBM). The effects of various parameters (such as confining

pressure and soil plasticity) on dynamic soil properties are evaluated and

quantified within this framework. One of the most important aspects of this study

is estimating not only the mean values of the empirical curves but also estimating

the uncertainty associated with these values. This study provides the opportunity

to handle uncertainty in the empirical estimates of dynamic soil properties within

the probabilistic seismic hazard analysis framework. A refinement in site-specific

probabilistic seismic hazard assessment is expected to materialize in the near

future by incorporating the results of this study into the state of practice.

viii
TABLE OF CONTENTS

LIST OF TABLES ...............................................................................................xiii

LIST OF FIGURES............................................................................................xviii

CHAPTER 1 INTRODUCTION........................................................................ 1
1.1 Background ........................................................................................... 1
1.2 Dynamic Soil Properties........................................................................ 4
1.3 Ground Response Analysis ................................................................... 8
1.4 Objectives of Research........................................................................ 10
1.5 Organization of Dissertation ............................................................... 11

CHAPTER 2 LABORATORY TESTING EQUIPMENT............................... 13


2.1 Introduction ......................................................................................... 13
2.2 Combined Resonant Column and Torsional Shear Equipment........... 14
2.3 Torsional Resonant Column Test ........................................................ 16
2.4 Cyclic Torsional Shear Test ................................................................ 21
2.5 Summary ............................................................................................. 22

CHAPTER 3 PHYSICAL PROPERTIES OF TEST SPECIMENS ................ 23


3.1 Introduction ......................................................................................... 23
3.2 Undisturbed Soil Specimens from Northern California...................... 25
3.3 Undisturbed Soil Specimens from Southern California...................... 29
3.4 Undisturbed Soil Specimens from South Carolina ............................. 35
3.5 Undisturbed Soil Specimens from Lotung, Taiwan ............................ 38
3.6 Overview of The Database .................................................................. 39
3.7 Summary ............................................................................................. 53

CHAPTER 4 OBSERVED TRENDS IN DYNAMIC SOIL PROPERTIES .. 54


4.1 Introduction ......................................................................................... 54
4.2 Background ......................................................................................... 54
4.3 Nonlinear Dynamic Soil Properties..................................................... 56

ix
4.4 Effect of Duration of Confinement on Small-Strain Dynamic Soil
Properties............................................................................................. 59
4.5 Effect of Effective Confining Pressure ............................................... 61
4.6 Effect of Overconsolidation Ratio....................................................... 70
4.7 Effect of Number of Cycles ................................................................ 74
4.8 Effect of Loading Frequency............................................................... 76
4.9 Effect of Soil Type .............................................................................. 81
4.10 Effect of Sample Disturbance ............................................................. 90
4.11 Summary ........................................................................................... 104

CHAPTER 5 EMPIRICAL RELATIONSHIPS ............................................ 107


5.1 Introduction ....................................................................................... 107
5.2 Hardin and Drnevich (1972) Design Equations ................................ 107
5.3 Empirical Relationships .................................................................... 113
5.4 Summary ........................................................................................... 129

CHAPTER 6 PROPOSED SOIL MODEL .................................................... 131


6.1 Introduction ....................................................................................... 131
6.2 Normalized Modulus Reduction Curve............................................. 132
6.3 Nonlinear Material Damping Curve.................................................. 134
6.4 Parametric Study of The Soil Model................................................. 147
6.5 Summary ........................................................................................... 152

CHAPTER 7 STATISTICAL ANALYSIS OF COLLECTED DATA


USING FIRST-ORDER, SECOND-MOMENT BAYESIAN METHOD 154
7.1 Introduction ....................................................................................... 154
7.2 Bayesian Approach ........................................................................... 155
7.3 First-Order, Second-Moment Bayesian Method ............................... 164
7.4 Form of Proposed Equations ............................................................. 172
7.5 Summary ........................................................................................... 179

x
CHAPTER 8 STATISTICAL ANALYSIS OF THE RCTS DATA.............. 180
8.1 Introduction ....................................................................................... 180
8.2 Analysis of Subsets of The Data ....................................................... 184
8.3 Analysis of All Credible Data ........................................................... 212
8.4 Summary ........................................................................................... 217

CHAPTER 9 PREDICTING NONLINEAR SOIL BEHAVIOR USING


THE CALIBRATED MODEL................................................................... 220
9.1 Introduction ....................................................................................... 220
9.2 Calculation of Reference Strain, Curvature Coefficient, Small-
Strain Material Damping Ratio and the Scaling Coefficient............. 221
9.3 Estimation of Normalized Modulus Reduction and Material
Damping Curves................................................................................ 224
9.4 Effect of Overconsolidation Ratio, Loading Frequency and
Number of Loading Cycles on Nonlinear Soil Behavior .................. 228
9.5 Effect of Confining Pressure on Nonlinear Soil Behavior................ 234
9.6 Effect of Soil Type on Nonlinear Soil Behavior ............................... 238
9.7 Effects of Confining Pressure and Soil Type on Stress-Strain
Curves................................................................................................ 242
9.8 Summary ........................................................................................... 248

CHAPTER 10 RECOMMENDED NORMALIZED MODULUS


REDUCTION AND MATERIAL DAMPING CURVES ......................... 249
10.1 Introduction ....................................................................................... 249
10.2 Effect of PI at a Given Mean Effective Stress .................................. 250
10.3 Effect of Mean Effective Stress on a Soil with Given Plasticity ...... 250
10.4 Impact of Utilizing the Recommended Curves on Earthquake
Response Predictions of Deep Sites .................................................. 250
10.5 Summary ........................................................................................... 272

CHAPTER 11 UNCERTAINTY ASSOCIATED WITH THE MODEL


PREDICTIONS .......................................................................................... 273
11.1 Introduction ....................................................................................... 273
11.2 Uncertainty in Nonlinear Soil Behavior............................................ 273

xi
11.3 Uncertainty in Predicted Ground Motions Due to the Uncertainty
in Nonlinear Soil Behavior................................................................ 284
11.4 Summary ........................................................................................... 295

CHAPTER 12 SUMMARY AND CONCLUSIONS....................................... 296


12.1 Summary ........................................................................................... 296
12.2 Conclusions ....................................................................................... 301

APPENDIX A ..................................................................................................... 303

APPENDIX B ..................................................................................................... 306

APPENDIX C ..................................................................................................... 311

APPENDIX D ..................................................................................................... 338

REFERENCES.................................................................................................... 357

VITA ................................................................................................................... 363

xii
LIST OF TABLES

Table 3.1 Physical properties of soils recovered from Oakland Outer


Harbor and test pressures (Hwang, 1997) ..................................... 24

Table 3.2 Physical properties of soils recovered from Treasure Island


and test pressures (Hwang and Stokoe, 1993b; and Hwang,
1997).............................................................................................. 25

Table 3.3 Physical properties of soils recovered from San Francisco


Airport and test pressures (Hwang, 1997)..................................... 27

Table 3.4 Physical properties of soils recovered from Gilroy and test
pressures (Hwang and Stokoe, 1993c; Hwang, 1997; and
Stokoe et al., 2001)........................................................................ 27

Table 3.5 Physical properties of soils recovered from Garner Valley


and test pressures (Stokoe and Darendeli, 1998) .......................... 28

Table 3.6 Physical properties of soils recovered from San Francisco-


Oakland Bay Bridge Site and test pressures (Stokoe et al.,
1998d)............................................................................................ 28

Table 3.7 Physical properties of soils recovered from Corralitos and


test pressures (Stokoe et al., 2001) ................................................ 28

Table 3.8 Physical properties of soils recovered from Borrego and test
pressures (Hwang, 1997)............................................................... 32

Table 3.9 Physical properties of soils recovered from Arleta and test
pressures (Darendeli and Stokoe, 1997; and Darendeli, 1997) ..... 32

Table 3.10 Physical properties of soils recovered from Kagel and test
pressures (Darendeli and Stokoe, 1997; and Darendeli, 1997) ..... 32

Table 3.11 Physical properties of soils recovered from La Cienega and


test pressures (Darendeli and Stokoe, 1997; Darendeli, 1997;
and Stokoe et al., 1998e) ............................................................... 33

Table 3.12 Physical properties of soils recovered from Newhall and test
pressures (Darendeli and Stokoe, 1997; and Darendeli, 1997) ..... 33

xiii
Table 3.13 Physical properties of soils recovered from Sepulveda V.A.
Hospital and test pressures (Darendeli and Stokoe, 1997; and
Darendeli, 1997)............................................................................ 34

Table 3.14 Physical properties of soils recovered from Potrero Canyon


and test pressures (Stokoe et al., 1998e) ....................................... 34

Table 3.15 Physical properties of soils recovered from Rinaldi Receiving


Station and test pressures (Stokoe et al., 1998e). .......................... 34

Table 3.16 Physical properties of soils recovered from North Palm


Springs and test pressures (Stokoe et al., 2001) ............................ 35

Table 3.17 Physical properties of soils recovered from Imperial Valley


College and test pressures (Stokoe et al., 2001)............................ 35

Table 3.18 Physical properties of soils recovered from Savannah River


Site and test pressures (Hwang, 1997; and Stokoe et al.,
1998a)............................................................................................ 37

Table 3.19 Physical properties of soils recovered from Daniel Island and
test pressures (Stokoe et al., 1998b).............................................. 37

Table 3.20 Physical properties of soils recovered from Lotung site and
test pressures (Hwang and Stokoe, 1993a; and Hwang, 1997) ..... 39

Table 3.21 Distribution of soil samples according to the sample depth in


each geographic region.................................................................. 41

Table 3.22 Distribution of collected according to the isotropic confining


pressure in each geographic region ............................................... 42

Table 3.23 Distribution of soil samples according to the Unified Soil


Classification System (USCS) designation and sample depth ...... 44

Table 4.1 Parameters that control nonlinear soil behavior and their
relative importance in terms of affecting normalized modulus
reduction and material damping curves based on general
trends observed during the course of this study .......................... 105

Table 5.1 Parameters that control nonlinear soil behavior and their
relative importance in terms of affecting shear modulus and
material damping (Hardin and Drnevich, 1972b) ....................... 108

xiv
Table 7.1 Prior information provided in the discrete example.................... 160

Table 7.2 Prior information regarding the model parameters in the


FSBM example............................................................................ 165

Table 7.3 Prior covariance structure of the model parameters in the


FSBM example............................................................................ 165

Table 7.4 Data used to calibrate the model parameters in the FSBM
example ....................................................................................... 166

Table 7.5 Comparison of the prior and posterior information regarding


the model parameters in the FSBM example .............................. 169

Table 7.6 Posterior covariance structure of the model parameters in the


FSBM example............................................................................ 170

Table 7.7 Posterior covariance structure of the model parameters in the


FSBM example............................................................................ 171

Table 8.1 Distribution of specimens with soil type and geographic


location ........................................................................................ 181

Table 8.2 Distribution of specimens by soil group and geographic


location ........................................................................................ 181

Table 8.3 Distribution of specimens with soil type and geographic


location for the updated database ................................................ 182

Table 8.4 Distribution of specimens by soil group and geographic


location for the updated database ................................................ 183

Table 8.5 Prior mean values and variances of the model parameters ......... 185

Table 8.6 Updated mean values and variances of the model parameters
for the soils from Northern California......................................... 186

Table 8.7 Updated mean values and variances of the model parameters
for the soils from Southern California......................................... 191

Table 8.8 Updated mean values and variances of the model parameters
for the soils from South Carolina ................................................ 194

xv
Table 8.9 Updated mean values and variances of the model parameters
for the South Carolina soil groups affected by change in the
contents of the database............................................................... 198

Table 8.10 Updated mean values and variances of the model parameters
for the soils from Lotung, Taiwan............................................... 200

Table 8.11 Updated mean values and variances of the model parameters
for the four soil groups ................................................................ 207

Table 8.12 Comparison of the prior and updated mean values and
variances of the model parameters for all the credible data ........ 214

Table 8.13 Covariance structure of the updated model parameters for all
the credible data .......................................................................... 218

Table 10.1 Effect of PI on normalized modulus reduction curve: σo’ =


0.25 atm....................................................................................... 252

Table 10.2 Effect of PI on material damping curve: σo’ = 0.25 atm............. 252

Table 10.3 Effect of PI on normalized modulus reduction curve: σo’ =


1.0 atm......................................................................................... 254

Table 10.4 Effect of PI on material damping curve: σo’ = 1.0 atm............... 254

Table 10.5 Effect of PI on normalized modulus reduction curve: σo’ =


4.0 atm......................................................................................... 256

Table 10.6 Effect of PI on material damping curve: σo’ = 4.0 atm............... 256

Table 10.7 Effect of PI on normalized modulus reduction curve: σo’ = 16


atm............................................................................................... 258

Table 10.8 Effect of PI on material damping curve: σo’ = 16 atm................ 258

Table 10.9 Effect of σo’ on normalized modulus reduction curve: PI = 0


%.................................................................................................. 260

Table 10.10 Effect of σo’ on material damping curve: PI = 0 % ................... 260

xvi
Table 10.11 Effect of σo’ on normalized modulus reduction curve: PI =
15 %............................................................................................. 262

Table 10.12 Effect of σo’ on material damping curve: PI = 15 % ................. 262

Table 10.13 Effect of σo’ on normalized modulus reduction curve: PI =


30 %............................................................................................. 264

Table 10.14 Effect of σo’ on material damping curve: PI = 30 % ................. 264

Table 10.15 Effect of σo’ on normalized modulus reduction curve: PI =


50 %............................................................................................. 266

Table 10.16 Effect of σo’ on material damping curve: PI = 50 % ................. 266

Table 10.17 Effect of σo’ on normalized modulus reduction curve: PI =


100 %........................................................................................... 268

Table 10.18 Effect of σo’ on material damping curve: PI = 100 % ............... 268

Table 11.1 Predicted mean values and standard deviations accounting for
uncertainty in the values of model parameters and variability
due to modeled uncertainty ......................................................... 275

Table 11.2 Predicted covariance structure accounting for uncertainty in


the values of model parameters and variability due to
modeled uncertainty .................................................................... 276

Table 11.3 Predicted mean values and standard deviations accounting


only for variability due to modeled uncertainty .......................... 277

Table 11.4 Predicted covariance structure accounting only for variability


due to modeled uncertainty ......................................................... 278

Table 12.1 Parameters that control nonlinear soil behavior and their
relative importance in terms of affecting normalized modulus
reduction and material damping curves based on general
trends observed during the course of this study .......................... 297

xvii
LIST OF FIGURES

Figure 1.1 Evaluation of ground motion at a geotechnical site based on


vertically propagating shear waves between the bedrock and
ground surface ................................................................................. 2

Figure 1.2 Fourier amplitude of (a) the ground motion as a result of (b)
the bedrock motion at the geotechnical site shown in Figure
1.1.................................................................................................... 3

Figure 1.3 Representation of a soil deposit in terms of dynamic soil


properties in geotechnical earthquake engineering ......................... 4

Figure 1.4 Nonlinear stress-strain curve of soils and variation of secant


shear modulus with shearing strain amplitude ................................ 5

Figure 1.5 Estimation of shear modulus and material damping ratio


during cyclic loading....................................................................... 6

Figure 1.6 (a) Nonlinear shear modulus and (b) normalized modulus
reduction curves .............................................................................. 7

Figure 1.7 Nonlinear material damping ratio curve.......................................... 7

Figure 1.8 Field curves representing nonlinear soil behavior........................... 9

Figure 2.1 Simplified diagram of the RCTS device (from Stokoe et al.,
1999).............................................................................................. 14

Figure 2.2 Simplified cross-sectional view of the confining system


(from Hwang, 1997) ...................................................................... 15

Figure 2.3 General Configuration of RCTS Equipment (after Hwang,


1997).............................................................................................. 17

Figure 2.4 Frequency response curve measured in the RC test (from


Stokoe et al., 1999)........................................................................ 18

Figure 2.5 Material damping measurement in the RC test using the half-
power bandwidth (from Stokoe et al., 1999)................................. 18

xviii
Figure 2.6 Material damping measurement in the RC test using the free-
vibration decay curve (from Stokoe et al., 1999).......................... 19

Figure 2.7 Calculation of shear modulus and material damping ratio in


the TS test...................................................................................... 21

Figure 3.1 Map of Northern California showing the locations of the


geotechnical sites in this area ........................................................ 26

Figure 3.2 Map of Southern California showing the locations of the


three geotechnical sites outside the Los Angeles area .................. 30

Figure 3.3 Map of Los Angeles showing the locations of the seven
geotechnical sites in this area ........................................................ 31

Figure 3.4 Map of South Carolina showing the locations of the


geotechnical sites in this area ........................................................ 36

Figure 3.5 Map of Taiwan showing the location of Lotung site .................... 38

Figure 3.6 Distribution of soil samples with geographic region .................... 40

Figure 3.7 Distribution of the number of geotechnical sites with


geographic region.......................................................................... 40

Figure 3.8 Distribution of soil samples according to the sample depth.......... 41

Figure 3.9 Distribution of confining pressures at which nonlinear


measurements were performed...................................................... 42

Figure 3.10 Distribution of soil samples according to soil type as


classified by the Unified Soil Classification System (USCS)....... 43

Figure 3.11 Distribution of soil samples according to soil plasticity in


terms of the plasticity index, PI..................................................... 44

Figure 3.12 Distribution of soil samples according to total unit weight .......... 46

Figure 3.13 Distribution of soil samples according to dry unit weight ............ 46

Figure 3.14 Distribution of soil samples according to water content ............... 47

Figure 3.15 Distribution of soil samples according to void ratio ..................... 47

xix
Figure 3.16 Variation of dry unit weight with depth of (a) fine grained
and (b) coarse grained soils included in this study........................ 48

Figure 3.17 Variation of water content with depth of (a) fine grained and
(b) coarse grained soils included in this study .............................. 49

Figure 3.18 Variation of void ratio with depth of (a) fine grained and (b)
coarse grained soils included in this study .................................... 50

Figure 3.19 Distribution of soil samples according to estimated


overconsolidation ratio.................................................................. 51

Figure 3.20 Variation of estimated overconsolidation ratio with depth of


(a) fine grained and (b) coarse grained soils included in this
study .............................................................................................. 52

Figure 4.1 Linear elastic, nonlinear elastic and plastic strain ranges on
(a) normalized modulus reduction and (b) material damping
curves ............................................................................................ 57

Figure 4.2 Variation of (a) low-amplitude shear modulus, (b) low-


amplitude material damping ratio, and (c) void ratio with
magnitude and duration of isotropic confining pressure............... 60

Figure 4.3 Variation of (a) low-amplitude shear modulus, (b) low-


amplitude material damping ratio, and (c) void ratio with
effective isotropic confining pressure ........................................... 62

Figure 4.4 The effect of confining pressure on the variation of (a) shear
modulus, (b) normalized shear modulus, and (c) material
damping ratio with shearing strain amplitude as measured in
the torsional resonant column ....................................................... 65

Figure 4.5 The effect of confining pressure on normalized modulus


reduction curve (a) for soils with moderate plasticity, and (b)
for non-plastic soils evaluated as part of the ROSRINE study
(after Stokoe et al., 1999) .............................................................. 67

Figure 4.6 The effect of confining pressure on (a) normalized modulus


reduction and (b) material damping curves of silty sands
evaluated as part of the ROSRINE study (after Darendeli et
al., 2001)........................................................................................ 68

xx
Figure 4.7 Impact on nonlinear site response of accounting for the effect
of confining pressure on dynamic soil properties (after
Darendeli et al., 2001) ................................................................... 70

Figure 4.8 The effect of overconsolidation ratio on the variation of (a)


shear modulus, (b) material damping ratio, and (c) void ratio
with effective isotropic confining pressure as measured in the
torsional resonant column ............................................................. 71

Figure 4.9 The effect of overconsolidation ratio on the variation of (a)


shear modulus, (b) normalized shear modulus, and (c)
material damping ratio with shearing strain amplitude as
measured in the torsional resonant column ................................... 72

Figure 4.10 The effect of number of loading cycles on the variation of (a)
shear modulus, (b) normalized shear modulus, and (c)
material damping ratio with shearing strain amplitude as
determined in the combined RCTS testing ................................... 75

Figure 4.11 The effect of loading frequency on (a) low-amplitude shear


modulus, and (b) low-amplitude material damping ratio as
determined in the combined RCTS testing ................................... 77

Figure 4.12 Comparison of the effect of loading frequency on low-


amplitude shear modulus and low-amplitude material
damping ratio (from Stokoe and Santamarina, 2000) ................... 78

Figure 4.13 The effect of loading frequency on the variation of (a) shear
modulus, (b) normalized shear modulus, and (c) material
damping ratio with shearing strain amplitude as determined
in the combined RCTS testing ...................................................... 80

Figure 4.14 The effect of soil type on the variation of (a) low-amplitude
shear modulus, and (b) low-amplitude material damping ratio
with effective isotropic confining pressure as determined in
the combined RCTS testing........................................................... 82

Figure 4.15 The effect of soil type on the variation of low-amplitude


shear modulus with loading frequency as determined in the
combined RCTS testing ................................................................ 84

xxi
Figure 4.16 The effect of soil type on the variation of low-amplitude
material damping ratio with loading frequency as determined
in the combined RCTS testing ...................................................... 85

Figure 4.17 The effect of soil type on the normalized modulus reduction
curve as measured in the torsional resonant column..................... 86

Figure 4.18 The effect of soil type on the material damping curve
determined at (a) N ~ 1000 cycles, (b) N = 1 cycle, and (c) N
= 10 cycles from combined RCTS testing .................................... 87

Figure 4.19 The effect of soil type on normalized modulus reduction and
material damping curves (after Stokoe et al., 1999) ..................... 88

Figure 4.20 Comparison of field and laboratory measurements of shear


wave velocity at the La Cienega site in the ROSRINE project..... 91

Figure 4.21 Variation of sampling disturbance expressed in terms of Vs,


lab/Vs, field
and Gmax, lab/Gmax, field with the in-situ shear wave
velocity .......................................................................................... 93

Figure 4.22 Comparison of laboratory and field measurements of small


strain material damping ratio (from Stokoe et al., 1999) .............. 95

Figure 4.23 Comparison of nonlinear soil properties back-calculated from


the free-field downhole accelerations with the laboratory
measurements (from Zeghal et al., 1995)...................................... 96

Figure 4.24 Comparison of the variation of (a) low-amplitude shear


modulus, (b) low-amplitude material damping ratio, and (c)
void ratio with effective isotropic confining pressure of intact
(undisturbed) and reconstituted (remolded) specimens ................ 99

Figure 4.25 Comparison of the variation of (a) shear modulus, (b)


normalized shear modulus, and (c) material damping ratio
with shearing strain of intact (undisturbed) and reconstituted
(remolded) specimens ................................................................. 100

Figure 4.26 Comparison of the variation of (a) shear modulus, (b)


normalized shear modulus, and (c) material damping ratio
with shearing strain measured using various equipment on
companion soil samples (from Stokoe et al., 1999) .................... 102

xxii
Figure 5.1 Hyperbolic soil model proposed by Hardin and Drnevich
(1972b) ........................................................................................ 110

Figure 5.2 The normalized modulus reduction and material damping


curves estimated based on the hyperbolic model ........................ 112

Figure 5.3 The effect of confining pressure on normalized modulus


reduction curve for Toyoura Sand (Iwasaki et al., 1978)............ 114

Figure 5.4 The effect of confining pressure on (a) normalized modulus


reduction, and (b) material damping curves for Toyoura Sand
(Kokusho, 1980).......................................................................... 115

Figure 5.5 The effect of confining pressure on (a) normalized modulus


reduction, and (b) material damping curves for non-plastic
soils (Ni, 1987) ............................................................................ 116

Figure 5.6 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Seed et al. (1986).......................... 118

Figure 5.7 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Sun et al. (1988) for soils with
plasticity ...................................................................................... 119

Figure 5.8 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Idriss (1990) ................................. 121

Figure 5.9 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Vucetic and Dobry (1991)............ 122

Figure 5.10 The effect of confining pressure on (a) normalized modulus


reduction, and (b) material damping curves for non-plastic
soils (Ishibashi and Zhang, 1993) ............................................... 124

Figure 5.11 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Ishibashi and Zhang (1993).......... 125

Figure 5.12 Variation in empirical (a) normalized modulus reduction, and


(b) material damping curves with depth (EPRI, 1993c).............. 127

Figure 5.13 Variation in empirical (a) normalized modulus reduction, and


(b) material damping curves with soil type (EPRI, 1993c)......... 128

xxiii
Figure 6.1 Normalized modulus reduction curve (of a silty sand at 1 atm
effective confining pressure) represented using a modified
hyperbolic model......................................................................... 133

Figure 6.2 Stress-strain curve (of a silty sand at 1 atm effective


confining pressure) estimated based on a modified reference
strain model ................................................................................. 135

Figure 6.3 Hysteresis loop estimated by modeling stress-strain reversals


for two-way cyclic loading according to Masing behavior......... 137

Figure 6.4 Calculation of damping ratio utilizing a hysteresis loop............. 138

Figure 6.5 Variations of c1, c2 and c3 with curvature coefficient, a.............. 141

Figure 6.6 Damping curve estimated based on Masing behavior................. 143

Figure 6.7 Effect of high-amplitude cycling on low-amplitude shear


modulus and material damping ratio (from Stokoe and
Lodde, 1978) ............................................................................... 144

Figure 6.8 Comparison of the variation in F with shearing strain for


different values of p..................................................................... 145

Figure 6.9 (a) Damping curve estimated based on Masing behavior, (b)
adjusted curve using the scaling coefficient, and (c) shifted
curve using the small-strain material damping ratio ................... 146

Figure 6.10 Effect of reference strain on (a) normalized modulus


reduction, (b) stress-strain, and (c) material damping curves ..... 148

Figure 6.11 Effect of the curvature coefficient on the normalized modulus


reduction curve ............................................................................ 149

Figure 6.12 Effect of the curvature coefficient on the stress-strain curve


(a) at small and intermediate strains, and (b) at high strains....... 149

Figure 6.13 Effect of the curvature coefficient on the material damping


curve ............................................................................................ 150

Figure 6.14 Effect of Dmin on the material damping curve............................. 151

Figure 6.15 The effect of scaling coefficient on material damping curve...... 152

xxiv
Figure 7.1 Prior probability mass function for the discrete example ........... 159

Figure 7.2 Posterior probability mass function for the discrete example ..... 161

Figure 7.3 Imaginary correlation between model parameters upon


updating prior information based on limited number of
observations................................................................................. 170

Figure 7.4 Variation of standard deviation with normalized shear


modulus ....................................................................................... 176

Figure 7.5 Standard deviation modeled for normalized modulus


reduction curve ............................................................................ 177

Figure 7.6 Variation of standard deviation with material damping ratio ..... 178

Figure 7.7 Standard deviation modeled for material damping curve ........... 178

Figure 8.1 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for
“clean” sands from Northern California...................................... 188

Figure 8.2 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for sands
with high fines content from Northern California....................... 188

Figure 8.3 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for silts
from Northern California ............................................................ 189

Figure 8.4 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for clays
from Northern California ............................................................ 189

Figure 8.5 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for
“clean” sands from Southern California...................................... 192

Figure 8.6 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for sands
with high fines content from Southern California....................... 192

xxv
Figure 8.7 Comparisons of the measured and predicted values of (a)
normalized modulus and (b) material damping ratio for silts
from Southern California ............................................................ 193

Figure 8.8 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for clays
from Southern California ............................................................ 193

Figure 8.9 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for
“clean” sands from South Carolina ............................................. 195

Figure 8.10 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for sands
with high fines content from South Carolina .............................. 195

Figure 8.11 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for silts
from South Carolina .................................................................... 196

Figure 8.12 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for clays
from South Carolina .................................................................... 196

Figure 8.13 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for sands
with high fines content from South Carolina (After
Discarding Specimens UT-39-G and UT-39-M) ........................ 199

Figure 8.14 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for clays
from South Carolina (After Discarding Specimens UT-39-O
and UT-39-S)............................................................................... 199

Figure 8.15 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for sands
with high fines content from Lotung, Taiwan............................. 201

Figure 8.16 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for silts
from Lotung, Taiwan................................................................... 201

xxvi
Figure 8.17 (a) Normalized modulus reduction and (b) material damping
curves estimated for a nonplastic silty sand using updated
mean values of model parameters calibrated at different
geographic locations.................................................................... 203

Figure 8.18 (a) Normalized modulus reduction and (b) material damping
curves estimated for a moderate plasticity silt using updated
mean values of model parameters calibrated at different
geographic locations.................................................................... 204

Figure 8.19 (a) Normalized modulus reduction and (b) material damping
curves estimated for a moderate plasticity clay using updated
mean values of model parameters calibrated at different
geographic locations.................................................................... 205

Figure 8.20 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for
“clean” sands ............................................................................... 208

Figure 8.21 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for sands
with high fines content ................................................................ 208

Figure 8.22 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for silts ..... 209

Figure 8.23 Comparisons of the measured and predicted values of (a)


normalized modulus and (b) material damping ratio for clays ... 209

Figure 8.24 (a) Normalized modulus reduction and (b) material damping
curves estimated using updated mean values of model
parameters calibrated for different soil groups ........................... 211

Figure 8.25 All credible (a) normalized modulus data from the resonant
column tests, and (b) material damping data from the
resonant column and torsional shear tests utilized to calibrate
the model parameters. ................................................................. 213

Figure 8.26 Comparisons of the measured and predicted values of


normalized modulus for all credible data .................................... 215

Figure 8.27 Comparisons of the measured and predicted values of


material damping for all credible data......................................... 216

xxvii
Figure 9.1 Estimation of reference strain for given values of PI, OCR
and in-situ mean effective stress ................................................. 223

Figure 9.2 Estimation of scaling coefficient for a given value of number


of loading cycles.......................................................................... 223

Figure 9.3 Estimation of small-strain material damping ratio for given


values of PI, OCR, in-situ mean effective stress and loading
frequency..................................................................................... 225

Figure 9.4 Estimated (a) normalized modulus reduction and (b) material
damping curves for the soil type and loading conditions
discussed in Section 9.2 .............................................................. 227

Figure 9.5 Effect of overconsolidation ratio on (a) normalized modulus


reduction and (b) material damping curves predicted by the
calibrated model .......................................................................... 229

Figure 9.6 Effect of loading frequency on (a) normalized modulus


reduction and (b) material damping curves predicted by the
calibrated model .......................................................................... 231

Figure 9.7 Effect of number of loading cycles on (a) normalized


modulus reduction and (b) material damping curves predicted
by the calibrated model ............................................................... 232

Figure 9.8 Comparison of (a) normalized modulus reduction and (b)


material damping curves predicted for resonant column and
torsional shear tests ..................................................................... 233

Figure 9.9 Effect of confining pressure on (a) normalized modulus


reduction and (b) material damping curves predicted by the
calibrated model .......................................................................... 235

Figure 9.10 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed for sands by Seed et al. (1986) .......... 236

Figure 9.11 Comparison of the effect of confining pressure on nonlinear


soil behavior of sand (PI = 0 %) predicted by the calibrated
model and empirical curves proposed for sands by Seed et al.
(1986) .......................................................................................... 237

xxviii
Figure 9.12 Effect of soil plasticity on (a) normalized modulus reduction
and (b) material damping curves predicted by the calibrated
model ........................................................................................... 239

Figure 9.13 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Vucetic and Dobry (1991)............ 240

Figure 9.14 Comparison of the effect of soil plasticity on nonlinear soil


behavior predicted by the calibrated model and empirical
curves proposed by Vucetic and Dobry (1991)........................... 241

Figure 9.15 Comparison of the measured in-situ shear wave velocities


and values predicted using Equation 9.4..................................... 244

Figure 9.16 Effect of confining pressure on stress-strain curve predicted


by the calibrated model for shearing strains ranging (a) from
γ = 0 to 1 % and (b) from γ = 0 to 0.01 % ................................... 245

Figure 9.17 Effect of soil plasticity on stress-strain curve predicted by the


calibrated model for shearing strains ranging (a) from γ = 0 to
1 % and (b) from γ = 0 to 0.01 % ................................................ 246

Figure 9.18 Comparison of the stress-strain curves of a sand and a


moderate plasticity clay based on the calibrated model for
shearing strains ranging (a) from γ = 0 to 1 % and (b) from γ
= 0 to 0.01 % ............................................................................... 247

Figure 10.1 Effect of PI on (a) normalized modulus reduction and (b)


material damping curves at 0.25 atm confining pressure............ 251

Figure 10.2 Effect of PI on (a) normalized modulus reduction and (b)


material damping curves at 1.0 atm confining pressure.............. 253

Figure 10.3 Effect of PI on (a) normalized modulus reduction and (b)


material damping curves at 4.0 atm confining pressure.............. 255

Figure 10.4 Effect of PI on (a) normalized modulus reduction and (b)


material damping curves at 16 atm confining pressure............... 257

Figure 10.5 Effect of mean effective stress on (a) normalized modulus


reduction and (b) material damping curves of a nonplastic
soil ............................................................................................... 259

xxix
Figure 10.6 Effect of mean effective stress on (a) normalized modulus
reduction and (b) material damping curves of a soil with PI =
15 %............................................................................................. 261

Figure 10.7 Effect of mean effective stress on (a) normalized modulus


reduction and (b) material damping curves of a soil with PI =
30 %............................................................................................. 263

Figure 10.8 Effect of mean effective stress on (a) normalized modulus


reduction and (b) material damping curves of a soil with PI =
50 %............................................................................................. 265

Figure 10.9 Effect of mean effective stress on (a) normalized modulus


reduction and (b) material damping curves of a soil with PI =
100 %........................................................................................... 267

Figure 10.10 Shear wave velocity profile assumed for the 100-m thick silty
sand deposit ................................................................................. 269

Figure 10.11 An example of utilizing the recommended normalized


modulus reduction and material damping curves and its
impact on estimated nonlinear site response ............................... 271

Figure 11.1 Mean values and standard deviations associated with the
point estimates of (a) normalized modulus reduction and (b)
material damping curves ............................................................. 280

Figure 11.2 Comparison of the correlated random realization of (a)


normalized modulus reduction and (b) material damping
curves relative to the mean curves and one standard deviation
ranges shown in Figure 11.1 ....................................................... 283

Figure 11.3 Comparison of spectral accelerations calculated using


perfectly correlated soil layers with µ, µ+σ and µ−σ
normalized modulus reduction and material damping curves..... 286

Figure 11.4 Comparison of spectral accelerations calculated using


perfectly correlated soil layers with 1) µ curves, 2) +σ
normalized modulus reduction and −σ material damping
curves, and 3) −σ normalized modulus reduction and
+σ material damping curves........................................................ 288

xxx
Figure 11.5 Fifty realizations of spectral acceleration computed using
completely uncorrelated soil layers with randomly generated
normalized modulus reduction and material damping curves..... 290

Figure 11.6 Histograms of spectral accelerations from fifty realizations


presented in Figure 11.5 (a) at 0.1 sec and (b) at 0.3 sec ............ 291

Figure 11.7 Histograms of spectral accelerations from fifty realizations


presented in Figure 11.5 (a) at 1 sec and (b) at 3 sec .................. 292

Figure 11.8 Distribution of fifty realizations of spectral acceleration


presented in Figure 11.5 .............................................................. 293

Figure 11.9 Comparison of the spectral accelerations from the fifty


realizations with the results computed utilizing mean
normalized modulus reduction and material damping curves..... 294

Figure 12.1 Comparison of the effect of confining pressure on nonlinear


soil behavior of sand (PI = 0 %) predicted by the calibrated
model and empirical curves proposed for sands by Seed et al.
(1986) .......................................................................................... 299

Figure 12.2 Comparison of the effect of soil plasticity on nonlinear soil


behavior predicted by the calibrated model and empirical
curves proposed by Vucetic and Dobry (1991)........................... 300

Figure 12.3 Mean values and standard deviations associated with the
point estimates of (a) normalized modulus reduction and (b)
material damping curves ............................................................. 302

xxxi
CHAPTER 1
INTRODUCTION

1.1 BACKGROUND

In earthquake engineering, the energy released during an earthquake is

represented by stress waves propagating through the bedrock and surfacing at the

site of interest. In terms of the geotechnical characteristics of the site, the site is

typically modeled as a series of horizontal layers with varying properties. In most

cases, the site is represented by softer soils close to the surface and stiffer soils at

depth. The increase in stiffness with depth is due to the older age of deeper

material and the confining effect of the overburden. Because of the progressive

increase in stiffness with depth, stress waves coming from depth often surface in a

propagation direction that is almost vertical.

Often times, an earthquake analysis includes predicting the dynamic

response of a structure at the geotechnical site. Since structures are always

designed with a factor of safety to support a static load (its self weight and the live

load) as a result of 1g vertical acceleration, the vertical component of the ground

motion does not generally have as much an impact on earthquake resistant design

as the horizontal component for which less precaution is often taken in the static

design.

With vertically propagating shear waves and a higher susceptibility of

structures to horizontal motions, the ground motion in many earthquake problems

is simply modeled as horizontal shaking due to vertically propagating shear

1
waves. In such a model, the soil deposit acts like a filter that amplifies energy at

some frequencies while attenuating it at others. Therefore, the estimated ground

motion is a function of the earthquake event and the local soil conditions as

shown in Figure 1.1. Two acceleration-time records are presented in this figure.

One of these is the bedrock motion and the second is the ground motion estimated

based on the bedrock motion and characteristics of the soil deposit.

üground
0.5
SOIL LAYER 1 Ground
Acceleration, 0.0
g
SOIL LAYER 2 -0.5
0 10 20 30 40 50 60

SOIL LAYER .. Time, sec

0.5
SOIL LAYER n Bedrock
Acceleration, 0.0
g
übedrock -0.5
BEDROCK 0 10 20 30 40 50 60
Time, sec

Figure 1.1 Evaluation of ground motion at a geotechnical site based on


vertically propagating shear waves between the bedrock and ground
surface

The filtering effect of the soil deposit is demonstrated in Figure 1.2 by

looking at the Fourier amplitude spectra of the two acceleration records. In this

figure, the acceleration components at different frequencies are shown for the

motions at the bedrock and ground surface. In this case, the low-frequency

motions (below 3 Hz) are amplified significantly. On the other hand, the high-
2
frequency motions are slightly attenuated. This effect can also be observed from

the comparison of the time records presented in Figure 1.1. Different cycles can

more easily be identified in the ground motion time record than in the bedrock

record.

0.010
(a)
0.008

Fourier 0.006
Amplitude,
g * sec 0.004

0.002

0.000

0.010
(b)
0.008

Fourier 0.006
Amplitude,
g * sec 0.004

0.002

0.000
0 2 4 6 8 10

Frequency, Hz

Figure 1.2 Fourier amplitude of (a) the ground motion as a result of (b) the
bedrock motion at the geotechnical site shown in Figure 1.1

3
1.2 DYNAMIC SOIL PROPERTIES

As discussed above, to analyze the response of structures during an

earthquake, it is necessary to characterize the ground motion underneath the

structure caused by the earthquake. Some of the most important ground motion

parameters are amplitude of motion (e.g., peak acceleration, peak velocity and

peak displacement), frequency content (e.g., Fourier spectra, response spectra,

predominant period, bandwidth) and duration. These parameters are primarily

affected by three factors: 1. source effects or the characteristics of the earthquake

(such as amount of energy released and type of faulting), 2. path effects (the

distance from the point of energy release to the site), and 3. site effects (such as

characteristics of the soil deposit, topography and other near-surface features).

This study focuses on characterization of the soil deposit. The properties that

typically need to be characterized are shear modulus, G, and material damping

ratio, D, as presented in Figure 1.3.


Shear
SOIL DEPOSIT Modulus, G Material
Damping
Ratio, D
BEDROCK

Figure 1.3 Representation of a soil deposit in terms of dynamic soil properties


in geotechnical earthquake engineering

4
Shear modulus, G, represents the shear stiffness of the soil. It is essentially

the slope of the relationship between shear stress (τ) and shearing strain (γ).

Because of the nonlinear nature of the stress-strain curve of soils, shear modulus

of soils change with strain amplitude as shown in Figure 1.4. The secant shear

modulus can also be approximated for the case of dynamic loading over a cycle of

loading at a given strain amplitude as shown in Figure 1.5. The stress-strain path

illustrated in this figure is called a hysteresis loop. The slope of the line that

connects the end points of the hysteresis loop represents the “average” shear

stiffness of the soil, hence the secant shear modulus.

Shear
Stress, τ

G2
G1 1
1

γ1 γ2 Shearing
Strain, γ

Figure 1.4 Nonlinear stress-strain curve of soils and variation of secant shear
modulus with shearing strain amplitude

5
Shear G
Stress, τ
1

AT

AL Shearing Strain, γ

G=τ/γ
D = AL / (4 π AT)

Figure 1.5 Estimation of shear modulus and material damping ratio during
cyclic loading

Material damping ratio, D, is a measure of the proportion of dissipated

energy to the maximum retained strain energy during each cycle at a given strain

amplitude as shown in Figure 1.5. The energy dissipated over a loading cycle is

represented by the gray area within the hysteresis loop (AL), and the maximum

retained strain energy is represented by the triangular area (AT) that is calculated

using peak shear stress and peak shearing strain. Material damping ratio is a result

of friction between soil particles, strain rate effects and nonlinearity of the stress-

strain relationship in soils.

As presented in Figure 1.4, soils exhibit nonlinear behavior in shear. The

secant shear modulus decreases with increasing strain amplitude as shown in

Figure 1.6a. Shear modulus at small strains, at which soil behavior is linear, is

referred to as small-strain shear modulus, Gmax. The relationship between shear

6
modulus and strain amplitude is typically characterized by a normalized modulus

reduction curve as shown in Figure 1.6b.

(a) (b)

120
1.0
80 G
G,
Gmax Gmax 0.5
MPa
40

0 0
0.001 0.01 0.1 1 0.001 0.01 0.1 1
Shearing Strain, γ , % Shearing Strain, γ , %

Figure 1.6 (a) Nonlinear shear modulus and (b) normalized modulus reduction
curves

The nonlinearity in the stress-strain relationship results in an increase in

energy dissipation and, therefore, an increase in material damping ratio with

increasing strain amplitude as presented in Figure 1.7. Material damping ratio at

small strains (in the linear range) is referred to as small-strain material damping

ratio, Dmin, herein.

16

D,
% 8 Dmin

0
0.001 0.01 0.1 1
Shearing Strain, γ , %

Figure 1.7 Nonlinear material damping ratio curve

7
1.3 GROUND RESPONSE ANALYSIS

In analyzing ground motions due to small vibrations, soil behavior is

assumed to be linear. Each soil layer is assigned a shear modulus and a material

damping ratio. Since a horizontally layered system is being modeled, the task of

ground response analysis is reduced to a simple 1-D wave propagation problem

that has a closed-form solution (Kramer, 1996).

On the other hand, dynamic soil properties can be extremely nonlinear

when ground motions are caused by large vibrations (such as design level

earthquakes). As a result, the change in shear modulus and material damping ratio

with shearing strain amplitude must be accounted for in ground response analysis.

The linear solution, which is applicable for small vibration levels, can be modified

to overcome this problem.

One approach to handling nonlinear soil behavior due to shaking during a

design level event is to perform linear analyses with dynamic soil properties that

are iterated in a manner consistent with an “effective” shearing strain induced in

the soil layer (Schnabel et al., 1972; and EduPro, 1998). This iterative approach is

called equivalent linear analysis.

The effective shearing strain is defined as a certain portion of the

maximum strain amplitude throughout the time history. The ratio of effective

shearing strain to maximum strain amplitude is typically related to the magnitude

of the earthquake event or the characteristics of the acceleration-time record

employed in the analysis. When a design level earthquake is analyzed, the ratio of

effective to maximum shearing strain is typically about 0.6.

8
The state of practice in equivalent linear analysis often employs empirical

normalized modulus reduction and a material damping curves. These empirical

curves are developed based on laboratory studies performed over the past three

decades.

The empirical normalized modulus reduction curve is scaled using an

estimate of the small-strain shear modulus, Gmax. The small-strain shear modulus

can be calculated using shear wave velocity, Vs, from in-situ seismic

measurements and mass density, ρ.

Gmax = ρ * Vs2 (1.1)

The curve calculated by scaling the empirical normalized modulus

reduction curve is called the field shear modulus curve (Figure 1.8). Since

material damping ratio can not be estimated accurately in-situ, the field material

damping curve is assumed to be identical to the empirical material damping curve

as shown in Figure 1.8.

Gmax, field
150 16

100
G, D,
MPa 8
%
50

0 0
0.001 0.01 0.1 1 0.001 0.01 0.1 1
Shearing Strain, γ , % Shearing Strain, γ , %

Gfield = Gmax, field * ( GG )


max
empirical
Dfield = Dempirical

Figure 1.8 Field curves representing nonlinear soil behavior


9
1.4 OBJECTIVES OF RESEARCH

As part of various research projects [including the SRS (Savannah River

Site) Project AA891070, EPRI (Electric Power Research Institute) Project 3302,

and ROSRINE (Resolution of Site Response Issues from the Northridge

Earthquake) Project] numerous sites were drilled and sampled. Intact soil samples

over a depth range of several hundred meters were recovered from 20 of these

sites. These soil samples were tested in the soil dynamics laboratory at The

University of Texas at Austin (UTA) to characterize the materials.

The presence of a database accumulated from testing these intact

specimens motivated a re-evaluation of empirical curves often employed in

seismic site response analyses. The weaknesses of empirical curves reported in

the literature were recognized and the necessity of developing an improved set of

empirical curves was acknowledged.

This study focuses on generating an improved set of empirical curves that

can be represented in the form of a set of simple equations. The data collected

over the past decade at The University of Texas at Austin are statistically

analyzed using the First-order, Second-moment Bayesian Method (FSBM). The

effects of various parameters (such as confining pressure and soil plasticity) on

dynamic soil properties are evaluated and quantified within this framework.

One of the most important aspects of this study is estimating not only the

mean values of the empirical curves but also the uncertainty associated with these

values. The handling of uncertainty in the empirical estimates of dynamic soil

10
properties is expected to result in a refinement of probabilistic seismic hazard

analysis.

1.5 ORGANIZATION OF DISSERTATION

A general overview of the dynamic laboratory test equipment used to

evaluate the nonlinear soil properties is presented in Chapter Two along with a

brief review of the theory upon which the laboratory testing is founded.
Information regarding the soil samples analyzed in this work is

summarized in Chapter Three. All testing was conducted at The University of

Texas at Austin over the past decade.

The sensitivity of dynamic soil properties to soil type and loading

conditions are described in Chapter Four. The general trends (in terms of how

these parameters affect nonlinear soil behavior) observed during the course of this

work and those reported in the literature are discussed.

The empirical relationships reported in the literature are summarized in

Chapter Five. The empirical normalized modulus reduction and material damping

curves proposed in the literature are evaluated in terms of capturing the general

trends discussed in Chapter Four.

A four-parameter soil model that describes the change in normalized shear

modulus and material damping ratio with shearing strain is presented in Chapter

Six along with a parametric study of the model. Two of these parameters,

reference strain and curvature coefficient, are utilized in describing the

normalized modulus reduction curve. Masing behavior is used as a criterion in

evaluating material damping. A scaling coefficient and small-strain material

11
damping ratio are utilized in describing the material damping curve relative to the

damping curve estimated from the normalized modulus reduction curve and

assuming Masing Behavior. The impact of soil type and loading conditions on the

model parameters are also described herein.

The First-order, Second-moment Bayesian method is briefly discussed in

Chapter Seven. The form of the equations that are used in relating model

parameters to soil type and loading conditions are discussed in this chapter.

Results of the statistical analysis are presented in Chapter Eight. Measured

and predicted curves are compared in order to evaluate the success of the model in

representing nonlinear soil behavior.

In Chapter Nine, the impact of soil type and loading conditions on model

parameters are quantified. Equations and graphical solutions that are utilized to

construct normalized shear modulus reduction and material damping curves for

different soil types and loading conditions are presented. These curves are

compared with other empirical curves reported in the literature.

In Chapter Ten, recommended normalized modulus reduction and material

damping curves are presented for soils with a broad range plasticity confined at

different mean effective stresses.

Uncertainty associated with the predicted normalized modulus reduction

and material damping curves is discussed in Chapter Eleven. Recommendations

for future work related with handling uncertainty in nonlinear soil behavior are

presented for probabilistic seismic hazard analysis.

A summary of the study and conclusions are presented in Chapter Twelve.

12
CHAPTER 2
LABORATORY TESTING EQUIPMENT

2.1 INTRODUCTION

Combined resonant column and torsional shear (RCTS) equipment was

employed in this work to evaluate the dynamic soil properties of undisturbed soil

specimens. This equipment was developed by Professor Stokoe and his graduate

students (Isenhower, 1979; Lodde, 1982; Ni, 1987; and Hwang, 1997) following

earlier designs by Hall and Richart (1963), Hardin and Music (1965), and

Drnevich (1967). Detailed information regarding the equipment, testing method,

theory and calibration is presented in Darendeli (1997).

The RCTS equipment uses a fixed-free configuration. The soil specimen

rests on a fixed bottom pedestal (fixed at the bottom) and is free at the top. At the

free end, four magnets are attached to the top cap and fixed coils surrounding the

magnets are used to excite the top of the specimen with torsional vibrations

without constraining the top of the specimen (hence the top of the specimen is

“free”). A simplified diagram of the combined RCTS equipment is presented in

Figure 2.1.

13
Resonant or Slow Cyclic
Torsional Excitation
Proximitor Probes Proximitor Target
Counter Weight Accelerometer

Drive M agnet
Coil Top Cap

Support
Plate
Rubber Securing
Specimen
M embrane Plate
Fluid Bath
Inner
Porous Cylinder
Stone O-ring
Base Plate

Figure 2.1 Simplified diagram of the RCTS device (from Stokoe et al., 1999)

2.2 COMBINED RESONANT COLUMN AND TORSIONAL SHEAR EQUIPMENT

Combined RCTS equipment is capable of testing a soil specimen in two

different modes. These modes are: 1. low frequency cyclic testing, and 2. higher

frequency dynamic testing during resonance. Thus, the same specimen can be

tested using both modes and variability due to testing different specimens or

testing the same specimen after it has been subjected to a different stress history is

eliminated. The data collected from the two independent modes of testing can

effectively be compared in order to gain more insight regarding material behavior.

One of the testing modes is called the torsional resonant column (RC) test,

which is based on the theory of torsional wave propagation in a fixed-free

cylinder with a mass attached at the free end. In this mode, well-defined boundary

14
conditions and specimen geometry are utilized in evaluating the shear modulus

and material damping ratio in shear from measurements at first-mode resonance.

The second testing mode is called the cyclic torsional shear (TS) test,

which involves monitoring the applied torque and displacement at the top of the

specimen. The torque is converted into shear stress and the displacement is

converted into shearing strain. Thus, hysteresis loops, which are utilized in

evaluation of shear modulus and material damping ratio, are generated.

These tests are typically carried out while the specimen is confined

isotropically. The confining chamber is designed to handle pressures up to 40

atmospheres (4.1 MPa). A cross-sectional view of the confining system is

presented in Figure 2.2.

Top Plate
Thin
Silicon Metal Tube
Fluid Bath Hollow
σ Cylinder

Fixing
Rod

Soil σ

Membrane
O-Ring
Air
Pressure
σ
Drainage

Figure 2.2 Simplified cross-sectional view of the confining system (from


Hwang, 1997)

15
The soil specimen is tested using both the cyclic torsional shear and

resonance modes simply by changing: 1) the amplitude and frequency of the

current in the drive coils, and 2) the motion monitoring devices (shown in Figure

2.3) used to record the specimen response. These changes are performed outside

the confining chamber; hence, they can be done without changing the state of

stress on the specimen.

2.3 TORSIONAL RESONANT COLUMN TEST

In torsional RC testing, a forcing function with fixed amplitude and

varying frequency is applied at the top of a cylindrical soil specimen. The output

from the accelerometer on the drive plate (shown in Figure 2.3) is recorded versus

the vibration frequency during a frequency sweep. The graph of accelerometer

output versus vibration frequency is called the frequency response curve. A

typical response curve is shown in Figure 2.4. The frequency at which the

accelerometer output reaches a maximum during first-mode torsional resonance is

denoted as the resonant frequency, fr, and it is used in calculating the shear wave

velocity of the specimen. The value of accelerometer output, Ar, at this frequency
is then used in calculating the peak shearing strain amplitude during the test.

The frequency response curve is also utilized in evaluating the material

damping ratio at small shearing strains, γ, (γ < 0.005 %). The half-power points

are identified as the two points on the frequency response curve with an amplitude

of 1/√2 times the peak value. The frequencies associated with the half-power

points, f1 and f2, are used in evaluating the material damping ratio as presented in
Figure 2.5.

16
(a) Top View
Magnet

Drive Plate
Counter
Weight
Support
Plate

A
Drive Coil
Holder
Accelerometer
A Proximitor Probe

(b) Section AA
Proximitor LVDT
Probe Accelerometer
Support Proximitor
Post Target Drive
Coil
Proximitor Magnet
Holder
Top Cap
Support
Leveling and Plate
Securing Screw
Securing
Specimen Plate
Porous Fluid Bath
Stone
Inner
Cylinder
Drainage Line
Base Pedestal

Figure 2.3 General Configuration of RCTS Equipment (after Hwang, 1997)

17
120

Accelerometer Output, mV
Resonance
I/Io=(ωrL/Vs) tan(ωrL/Vs)
Ar G = ρVs2
80 Ar →γ

40

fr = ωr / 2 π

0
35 40 45 50 55 60
Frequency, f, Hz

Figure 2.4 Frequency response curve measured in the RC test (from Stokoe et
al., 1999)

Figure 2.5 Material damping measurement in the RC test using the half-power
bandwidth (from Stokoe et al., 1999)

18
Once the resonant frequency is identified, a second measurement of

material damping ratio can be performed using the free-vibration decay curve.

This method involves vibrating the specimen in steady-state, first-mode torsional

resonance and recording the decay of free vibrations after shutting off the driving

force. Figure 2.6 shows an example decay curve.

3
,%

(a)
1 Cycle
-3

2 Number
Shearing Strain Amplitude, γ x 10

5
10
1 15
0

-1

-2
Steady State Free Vibration Decay
-3
0.0 0.1 0.2 0.3 0.4
Time, seconds
Normalized Peak-to-Peak Amplitudes

1.0 Remolded Sand


eo = 0.71
δ = 0.0734
0.7 D = 1.17 %

0.5

(b)
0.3
0 5 10 15 20
Number of Cycles

Figure 2.6 Material damping measurement in the RC test using the free-
vibration decay curve (from Stokoe et al., 1999)

19
The logarithmic decrement, δ, is defined from the free-vibration decay

curve as:
1  z1 
δ = ln  (2.1)
n  z n +1 

where n equals number of cycles between two peak points in the time record, and
z1 and zn+1 are the amplitudes of cycle 1 and cycle n+1, respectively (Richart, Hall

and Woods, 1970). Material damping ratio can then be calculated using Equation

2.2.
δ2
D= (2.2)
4Π 2 + δ 2

The half-power bandwidth method is based on the theory of elasticity and

it is accurate during testing at small strains as noted above (γ < 0.005 %). Material

damping estimates based on this method are quite reproducible at small strains

since points around the peak output on the frequency response curve are utilized

in the calculations. On the other hand, background noise can have a more adverse

effect on the free-vibration decay curve.

At large strains, nonlinear behavior of soil results in the linear assumption,

on which the half-power bandwidth method is based, to become invalid. In this

case, the free-vibration decay curve is applied along with an adjustment of the

strain amplitude, which is constantly changing as the vibrations decay. In the free-

vibration decay method, material damping ratio is calculated using the first three

cycles of vibration in this study. As a result, the average amplitude of the first

three cycles of vibration is assumed to represent the shearing strain at which the

measurement is made (rather than the steady-state amplitude).

20
2.4 CYCLIC TORSIONAL SHEAR TEST

In cyclic TS testing, a slow torsional loading is applied at the top of the

specimen. The loading frequency used in TS testing is much lower than resonance

testing (at least 10 times less than the resonant frequency). The current in the

calibrated drive coils is monitored, and the torque applied to the specimen is

calculated. The displacement at the top of the specimen is also monitored using

proximitors. Based on the torque and displacement at the top of the specimen,

hysteresis loops are generated.

The secant shear modulus for each cycle of loading is evaluated by

calculating the slope of the line that connects the end points of the hysteresis loop

as illustrated in Figure 2.7.

Shear G
Stress, τ
1

AT

AL Shearing Strain, γ

G=τ/γ
D = AL / (4 π AT)

Figure 2.7 Calculation of shear modulus and material damping ratio in the TS
test

21
Material damping ratio is evaluated by calculating the ratio of the area

within the hysteresis loop (AL) and the maximum potential energy stored in each

cycle of motion as represented by the triangular area (AT). The area AT is

calculated using the end point of the hysteresis loop as shown in Figure 2.7.
AL
D= (2.3)
4ΠAT

2.5 SUMMARY

Information regarding the RCTS equipment is summarized in this chapter.

This equipment has been employed in evaluating nonlinear dynamic soil

properties of undisturbed soil specimens for more than two decades at The

University of Texas at Austin. Detailed information regarding the equipment,

testing method, theory and calibration is presented in Ni (1987), Hwang (1997)

and Darendeli (1997).

22
CHAPTER 3
PHYSICAL PROPERTIES OF TEST SPECIMENS

3.1 INTRODUCTION

Over the past decade, a total of 110 undisturbed soil samples, which were

taken from 20 geotechnical sites, were tested in the soil dynamics laboratory

using the combined RCTS equipment. In this study, dynamic properties of these

soils in the nonlinear range are analyzed. Information regarding the soil samples

and the confining pressures at which these samples were tested are tabulated in

this chapter. Also, the geotechnical reports that contain the original data are cited

herein.

The chapter has been divided into the following sections. Sections 3.2

through 3.5 describe samples taken from Northern California, Southern

California, South Carolina and Lotung, Taiwan, respectively. In each section,

information regarding specimens from each geotechnical site in a given

geographic region is presented in a separate table. As an example, Table 3.1

shows physical properties of soils recovered from Oakland Outer Harbor. The

publication that contains the original data is cited in the title of the table. The table

provides the following information on each specimen: 1) Specimen identification

(ID), 2) Depth of the soil sample, 3) Soil type determined according to the Unified

Soil Classification System (USCS) based on gradation and plasticity tests

performed at The University of Texas at Austin, 4) Percentage of fine material

(passing #200 sieve) by weight (listed as Fines Content), 5) Liquid limit (LL), 6)

23
Plasticity index (PI) which is equal to the difference between the liquid limit and

the plastic limit of the soil sample, 7) Water content of the specimen, 8) Total unit

weight of the specimen, 9) Void ratio of the specimen, 10) Estimated

overconsolidation ratio (Est. OCR) of the specimen based on characteristics of the

measured relationship between small-strain shear modulus and mean effective

confining pressure, and 11) Mean effective confining pressure (listed as Test

Pressure) at which data regarding nonlinear soil behavior were collected. These

11 items represent the 11 column headings in the tables.

Table 3.1 Physical properties of soils recovered from Oakland Outer Harbor
and test pressures (Hwang, 1997)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UT-33-A 34 CL 73 49 29 41 1.78 1.15 1 2.3
UT-33-B 21 SM-SC 37 21 4 15 1.78 0.74 1 1.5
UT-33-C 5 SP 4 NP NP 20 1.75 0.83 2 0.3
UT-33-D 81 CH 100 62 37 15 1.67 0.87 1 5.4
UT-33-E 144 SM-SC 40 25 7 18 2.05 0.52 1 8.2
* Research was funded by EPRI.

Section 3.6 contains a discussion on the distribution of samples in terms of


their geographic location, depth, soil type, plasticity index, void ratio and unit

weight. This discussion is an attempt to familiarize the reader with the

characteristics of the database that is utilized in generating a new set of empirical

curves and equations regarding nonlinear soil behavior. Knowing the contents of

the database that this study has utilized, the reader will be aware when an

application requires extrapolation of these empirical curves and equations so that

the results will be used with more caution under such circumstances.

24
3.2 UNDISTURBED SOIL SPECIMENS FROM NORTHERN CALIFORNIA

A total of 37 undisturbed soil samples from 7 sites in Northern California

tested as part of a number of research projects are included in this study. These

projects were funded by the Kajima Corporation, Geovision, Agbabian

Associates, EPRI (Electric Power Research Institute), Fugro, Inc., and Earth

Mechanics, Inc.

The geotechnical sites in Northern California are Corralitos, Garner

Valley, Gilroy, Oakland Outer Harbor, San Francisco Airport, San Francisco-

Oakland Bay Bridge and Treasure Island. A map of Northern California is

presented in Figure 3.1 showing the locations of these sites. Information regarding

the soil samples and the confining pressures at which nonlinear properties were

measured are tabulated in Tables 3.1 through 3.7. The references that contain the

original data are also cited in the titles of the tables.

Table 3.2 Physical properties of soils recovered from Treasure Island and test
pressures (Hwang and Stokoe, 1993b; and Hwang, 1997)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UT-28-A 18 CH 50 51 26 50 1.73 1.34 1 1.2
UT-28-B 52 CL 79 34 19 21 2.05 0.58 1 3.8
UT-28-C 71 CL 67 48 30 33 1.84 0.95 1 5.1
UT-28-D 40 CL 63 37 23 37 1.83 1.02 1 2.9
UT-28-E 9.1 SP-SM 79 NP NP 21 1.92 0.67 1 0.7
UT-28-F 27 CL 58 42 19 42 1.81 1.10 1 1.9
UT-28-G 5.3 SM 80 NP NP 20 1.92 0.69 1 0.4
UT-28-H 34 SP-SM 78 NP NP 22 1.83 0.76 1 2.2
* Research was funded by EPRI.

25
Figure 3.1 Map of Northern California showing the locations of the
geotechnical sites in this area

26
Table 3.3 Physical properties of soils recovered from San Francisco Airport
and test pressures (Hwang, 1997)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UT-36-A 4.9 CL-ML 53 21 5 19 2.08 0.54 2 0.3
UT-36-B 7.9 CL 54 30 13 18 2.10 0.50 2 0.5
* Research was funded by EPRI.

Table 3.4 Physical properties of soils recovered from Gilroy and test pressures
(Hwang and Stokoe, 1993c; Hwang, 1997; and Stokoe et al., 2001)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UT-24-A 6.1 CL 86 43 23 30 1.91 0.84 1 0.8
UT-24-B 3.0 CL-ML 65 29 7 26 1.88 0.81 1 0.4
UT-24-C 26 MH 100 47 17 31 1.94 0.82 1 2.4
UT-24-D 37 ML 62 NP NP 20 2.12 0.48 1 3.3
++
UT-24-E 128 * * * * 14 2.18 0.41 1 8.7
UT-24-F 64 SW-SM 8 NP NP 15 2.08 0.46 1 4.9
UT-24-G 15 SP 2 NP NP 16 1.97 0.55 1 2.0
UT-24-H 106 CL 65 35 13 24 2.04 0.63 1 8.6
UT-24-I 6.1 CL 86 43 23 30 1.91 0.84 1 0.8
UT-24-J 26 MH 100 47 17 31 1.94 0.82 1 2.4
UT-24-K 37 ML 62 NP NP 20 2.12 0.48 1 3.3
++
UT-24-L 52 SM 13 NP NP 8 2.13 0.34 1 4.8
UTA-18-I 3.4 SC 17 36 20 20 2.15 0.47 1 0.4
UTA-18-J 16 SC 28 49 24 15 1.95 0.56 1 1.1
* Information is not available.
** Research was funded by EPRI and Kajima Corporation, Japan through Geovision.
++
UT-24-E and UT-24-L were stopped due to membrane leakage and are not included in this
study.

27
Table 3.5 Physical properties of soils recovered from Garner Valley and test
pressures (Stokoe and Darendeli, 1998)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UT-52-A 3.5 SM 26 NP NP 19 1.90 0.69 1 0.5
UT-52-B 6.5 SM 15 NP NP 17 1.79 0.76 1 0.7
UT-52-C 41 SM 36 NP NP 18 1.91 0.67 1 2.7
UT-52-D 27 SM 19 NP NP 14 2.09 0.47 1 1.7
* Research was funded by Agbabian Associates.

Table 3.6 Physical properties of soils recovered from San Francisco-Oakland


Bay Bridge Site and test pressures (Stokoe et al., 1998d)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-10-A 8.4 CH 93 63 36 50 1.71 1.38 1 0.5, 2.2
UTA-10-B 11 CH 97 75 43 58 1.63 1.61 1 0.5, 2.2
UTA-10-C 24 CH 96 89 53 57 1.70 1.50 8 1.1, 4.4
UTA-10-D 71 CL 91 46 19 30 1.92 0.83 1 4.1, 16.3
* Research was funded by Fugro, Inc. and Earth Mechanics, Inc.

Table 3.7 Physical properties of soils recovered from Corralitos and test
pressures (Stokoe et al., 2001)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-18-F 10 SW-SC 7 31 10 14 2.16 0.44 1 1.4
UTA-18-G** 3.3 SC 31 44 19 19 2.01 0.62 1 0.5
UTA-18-K 46 ML 52 23 4 8 2.39 0.24 2 4.1
* Research was funded by Kajima Corporation, Japan through Geovision.
** UTA-18-G was reconstituted and is not included in this study.

28
Specimen UTA-18-G sampled from Corralitos was reconstituted.

Specimens UT-24-E and UT-24-L from Gilroy had to be stopped due to

membrane leakage before the TS data could be collected. As a result, these

samples are not included in this study, but are included in the tables for

completeness.

3.3 UNDISTURBED SOIL SPECIMENS FROM SOUTHERN CALIFORNIA

A total of 47 undisturbed soil samples from 10 sites in Southern California

tested as part of a number of research projects are included in this study. These

projects were funded by the ROSRINE (Resolution of Site Response Issues from

the Northridge Earthquake) Project, Kajima Corporation, Geovision and

Agbabian Associates.

The geotechnical sites in Southern California are Arleta, Borrego, Imperial

Valley College, Kagel, La Cienega, Newhall, North Palm Springs, Potrero

Canyon, Rinaldi Receiving Station and Sepulveda V.A. Hospital. A map of

Southern California is presented in Figure 3.2 showing the locations of the three

sites outside the Los Angeles area (Borrego, Imperial Valley College and North

Palm Springs). The remaining seven sites are presented on a map of Los Angeles

area in Figure 3.3.

Information regarding the soil samples and the confining pressures (at

which these samples were tested) are tabulated in Tables 3.8 through 3.17. The

references that contain the original data are also cited in the titles of the tables.

29
Figure 3.2 Map of Southern California showing the locations of the three
geotechnical sites outside the Los Angeles area

Specimen UT-40-G from Borrego was stopped due to membrane leakage

before the TS data could be collected. Specimens UTA-9-I from Rinaldi

Receiving Station, and UTA-18-B and UTA-18-E from North Palm Springs were

reconstituted. As a result, these samples are not included in this study, but are

included in the tables for completeness.

30
Figure 3.3 Map of Los Angeles showing the locations of the seven geotechnical
sites in this area

31
Table 3.8 Physical properties of soils recovered from Borrego and test
pressures (Hwang, 1997)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UT-40-B 3.4 SM 84 NP NP 12 1.76 0.68 1 0.4
UT-40-C 20 SP-SM 89 NP NP 11 1.83 0.62 1 2.5
UT-40-E 49 SP-SM 91 NP NP 14 1.76 0.72 1 6.1
UT-40-F 110 SP-SM 91 NP NP 17 1.78 0.75 1 13.6
UT-40-G** 146 SP-SM 95 NP NP 11 2.04 0.44 1 18.2
* Research was funded by Agbabian Associates.
** UT-40-G was stopped due to membrane leakage and is not included in this study.

Table 3.9 Physical properties of soils recovered from Arleta and test pressures
(Darendeli and Stokoe, 1997; and Darendeli, 1997)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-1-C 31 SM 40 21 1 13 2.16 0.42 2 2.7, 10.9
UTA-1-L 15 SM 28 NP NP 14 2.10 0.46 1 1.6
* Research was funded by ROSRINE Project.

Table 3.10 Physical properties of soils recovered from Kagel and test pressures
(Darendeli and Stokoe, 1997; and Darendeli, 1997)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-1-E 8.5 SW-SM 10 NP NP 3 1.87 0.48 1 1.1, 4.4
UTA-1-F 31 SW-SM 12 NP NP 13 2.11 0.44 1 3.3
UTA-1-G 65 SW-SM 9 NP NP 10 2.16 0.38 1 5.4
UTA-1-H 92 SP-SM 10 NP NP 13 2.07 0.46 1 6.8
* Research was funded by ROSRINE Project.

32
Table 3.11 Physical properties of soils recovered from La Cienega and test
pressures (Darendeli and Stokoe, 1997; Darendeli, 1997; and Stokoe
et al., 1998e)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-1-J 4.9 CL 51 33 10 21 2.03 0.61 1 0.6, 2.4
UTA-1-K 3.4 SC 35 42 20 15 1.98 0.57 2 0.5, 2.2
UTA-1-M 7.9 SM 43 NP NP 22 2.07 0.58 1 1.1, 4.4
UTA-1-N 6.1 CL 60 29 8 19 1.99 0.62 2 0.8, 3.3
UTA-1-O 6.4 CL 57 32 10 20 2.05 0.58 1 0.8, 3.3
UTA-9-J 28 CH 99 50 25 30 1.87 0.87 4 3.3
UTA-9-K 34 CL 83 26 10 15 2.01 0.55 1 4.6
UTA-9-L 36 CL 64 30 10 20 2.08 0.56 1 4.6
UTA-9-M 95 SM 17 NP NP 16 2.07 0.51 1 13.6
UTA-9-N 125 SM 30 NP NP 16 2.10 0.49 1 17.0
UTA-9-O 186 ML 71 28 5 29 2.05 0.69 1 24.5
UTA-9-P 241 SM 32 NP NP 18 2.05 0.55 1 27.2
UTA-9-Q 52 CL 92 34 11 29 1.90 0.83 1 6.8
UTA-9-R 107 CH 99 64 36 32 1.96 0.82 1 13.6
UTA-9-S 150 CH 99 52 25 26 2.02 0.69 1 20.4
UTA-9-T 218 CL 94 42 18 21 2.04 0.60 1 27.2
* Research was funded by ROSRINE Project.

Table 3.12 Physical properties of soils recovered from Newhall and test
pressures (Darendeli and Stokoe, 1997; and Darendeli, 1997)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-1-D 62 SC-SM 48 25 5 13 2.21 0.37 4 5.4
UTA-1-I 21 SM 26 NP NP 18 1.92 0.65 1 1.6
* Research was funded by ROSRINE Project.

33
Table 3.13 Physical properties of soils recovered from Sepulveda V.A. Hospital
and test pressures (Darendeli and Stokoe, 1997; and Darendeli,
1997)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-1-A 3.4 CL 75 39 15 19 2.04 0.58 8 0.7
UTA-1-B 3.1 CL 62 37 15 20 1.91 0.70 8 0.5
UTA-1-P 14 ML 85 34 9 26 1.93 0.77 1 1.6
UTA-1-Q 17 SM 41 NP NP 15 2.00 0.55 1 2.2
UTA-1-R 37 CL 64 29 9 18 2.07 0.54 1 3.4
UTA-1-S 59 CL 66 42 16 22 2.14 0.54 2 5.4
UTA-1-T 2.4 CH 70 54 29 25 2.02 0.67 8 0.3, 1.4
UTA-1-U 86 CL 77 35 12 17 2.11 0.50 1 6.8
* Research was funded by ROSRINE Project.

Table 3.14 Physical properties of soils recovered from Potrero Canyon and test
pressures (Stokoe et al., 1998e)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-9-C 8.5 SC-SM 20 19 5 11 2.15 0.39 1 1.1
UTA-9-D 16 CL 73 26 9 16 2.13 0.47 1 1.9
UTA-9-E 31 CL 83 32 12 10 2.32 0.28 1 4.4
UTA-9-G 2.4 SM 41 20 2 21 1.88 0.74 1 0.4
* Research was funded by ROSRINE Project.

Table 3.15 Physical properties of soils recovered from Rinaldi Receiving


Station and test pressures (Stokoe et al., 1998e).

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-9-A 11 SM 22 23 1 15 2.04 0.51 1 1.4
UTA-9-B 21 CL-ML 53 23 5 24 1.99 0.69 1 2.4
UTA-9-F 2.4 CL-ML 51 22 4 22 2.03 0.62 1 0.4
UTA-9-H 15 SM 41 NP NP 16 2.03 0.55 1 1.6
UTA-9-I** 7.6 SW-SM 9 NP NP 12 2.10 0.44 1 1.1
* Research was funded by ROSRINE Project.
** UTA-9-I was reconstituted and is not included in this study.

34
Table 3.16 Physical properties of soils recovered from North Palm Springs and
test pressures (Stokoe et al., 2001)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-18-B** 72 SP 0 NP NP 16 1.93 0.59 1 4.8
UTA-18-C 46 SM 48 29 3 23 1.99 0.64 1 3.1
UTA-18-E** 17 SW 2 NP NP 18 2.17 0.44 1 1.4
* Research was funded by Kajima Corporation, Japan through Geovision.
** UTA-18-B and UTA-18-E were reconstituted and are not included in this study.

Table 3.17 Physical properties of soils recovered from Imperial Valley College
and test pressures (Stokoe et al., 2001)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-18-A 102 CL 100 46 29 21 1.93 0.69 1 6.5
UTA-18-H 16 CL 95 49 28 26 2.08 0.64 8 1.1
* Research was funded by Kajima Corporation, Japan through Geovision.

3.4 UNDISTURBED SOIL SPECIMENS FROM SOUTH CAROLINA

A total of 18 undisturbed soil samples from 2 sites in South Carolina were

tested as part of two research projects funded by the Westinghouse Savannah

River Company and S&ME, Inc.


The geotechnical sites in South Carolina are Savannah River Site and

Daniel Island. A map of South Carolina is presented in Figure 3.4 showing the

locations of these sites. Information regarding the soil samples and the confining

pressures at which nonlinear properties were measured are tabulated in Tables

3.18 and 3.19. The references that contain the original data are also cited in the

titles of the tables.

35
Figure 3.4 Map of South Carolina showing the locations of the geotechnical
sites in this area

36
Table 3.18 Physical properties of soils recovered from Savannah River Site and
test pressures (Hwang, 1997; and Stokoe et al., 1998a).

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UT-39-A 3.2 SC 30 52 31 15 2.07 0.48 4 0.8
UT-39-B 17 SM 20 NP NP 33 1.83 0.92 1 2.1
UT-39-C 27 CH 63 80 53 53 1.62 1.55 1 2.8
UT-39-D 7.0 SC 28 46 19 21 1.84 0.74 2 0.9
UT-39-E 47 SP-SM 9 NP NP 26 1.94 0.73 1 4.0
UT-39-F 57 SP-SM 11 NP NP 24 1.94 0.69 1 4.6
UT-39-G 86 SM 24 NP NP 20 1.65 0.93 1 6.3
UT-39-H 80 SM 20 NP NP 24 1.78 0.86 1 5.9
UT-39-I 24 SC 23 61 34 31 1.83 0.9 1 2.4
UT-39-K 13 SM 14 NP NP 27 1.81 0.85 1 1.6
UT-39-L 32 SM 18 NP NP 28 1.83 0.86 1 3.1
UT-39-M 263 SC 29 34 16 12 2.08 0.43 1 16.8
UT-39-N 107 CH 87 51 27 21 2.02 0.61 1 7.6
UT-39-O 226 CL 76 30 12 7 2.07 0.37 1 14.6
UT-39-S 199 CL 70 39 14 16 2.12 0.45 1 13.0
* Research was funded by Westinghouse Savannah River Company.
** As discussed in Chapter Eight, specimens UT-39-G, UT-39-M, UT-39-O, and UT-39-S were
removed from the database during the analysis because the resonant column results did not follow
the general trends reported in the literature and observed during the course of this study while the
torsional shear results did follow the general trends but were not of sufficient strain range to be
included.

Table 3.19 Physical properties of soils recovered from Daniel Island and test
pressures (Stokoe et al., 1998b).

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UTA-7-A 11 CH 93 122 79 84 1.51 2.31 1 0.4
UTA-7-B 10 SP-SM 8 26 2 38 1.84 1.03 1 0.6
UTA-7-C 20 MH 68 210 132 83 1.48 2.33 4 0.7, 2.7
* Research was funded by S&ME, Inc.

37
3.5 UNDISTURBED SOIL SPECIMENS FROM LOTUNG, TAIWAN

Eight samples from Lotung site in Taiwan were tested as part of a research

project funded by EPRI. Detailed information about this work can be found in

Hwang and Stokoe (1993a), and Hwang (1997).

A map of Taiwan is presented in Figure 3.5 showing the location of

Lotung. Information regarding the soil samples and the confining pressures at

which these samples were tested are tabulated in Table 3.20.

Lotung
8 Samples

Figure 3.5 Map of Taiwan showing the location of Lotung site

38
Table 3.20 Physical properties of soils recovered from Lotung site and test
pressures (Hwang and Stokoe, 1993a; and Hwang, 1997)

Fines Water Total Void Est. Test


Specimen Depth Soil Type Content LL PI Content Unit Wt. Ratio, OCR Pressure
3
ID (m) (USCS) (%) (%) (%) (%) (gr/cm ) e (atm)
UT-37-A 34 ML 99 32 7 35 1.89 0.92 1 2.0
UT-37-B 18 SM 40 NP NP 33 1.75 1.08 1 1.1
UT-37-C 5.5 ML 52 NP NP 31 1.79 0.97 1 0.3
UT-37-D 11 ML 85 NP NP 33 1.89 0.91 1 0.7
UT-37-E 29 SM 30 NP NP 31 1.91 0.88 1 1.7
UT-37-F 41 ML 98 33 8 31 1.88 0.93 1 2.4
UT-37-G 45 ML 78 NP NP 24 2.05 0.64 1 2.7
UT-37-H 25.0 ML 98 38 12 37 1.88 1.00 1 1.5
* Research was funded by EPRI.

3.6 OVERVIEW OF THE DATABASE

In this section, the distribution of the specimens included in this study is

briefly discussed in terms of their various characteristics. Figure 3.6 shows the

number of samples taken from each geographic region (Northern California,

Southern California, South Carolina and Taiwan). In Figure 3.7, the number of

geotechnical sites in each of the four geographic regions is presented. It is

important to note that most of the samples in this database have come from

California (84 out of 110 samples or 76 %).

Figure 3.8 shows the distribution of soil samples with depth. The samples

in this database have been recovered from a depth range of 3 to 263 m. This depth

range has been divided into eight categories as noted in the legend. The number of

samples from each geographic region in each depth category is presented in Table

3.21.

39
Figure 3.6 Distribution of soil samples with geographic region

Figure 3.7 Distribution of the number of geotechnical sites with geographic


region

40
Figure 3.8 Distribution of soil samples according to the sample depth

Table 3.21 Distribution of soil samples according to the sample depth in each
geographic region
Depth Range (meters)
Geographic Region 0-5 5-10 10-20 20-30 30-50 50-100 100-200 200-263 TOTAL
Northern California 5 8 4 6 7 5 2 - 37
Southern California 8 5 7 4 8 7 6 2 47
South Carolina 1 1 4 3 2 3 2 2 18
Taiwan - 1 2 2 3 - - - 8
TOTAL 14 15 17 15 20 15 10 4 110

In Figure 3.9, information regarding the isotropic confining pressures at

which nonlinear measurements have been performed is presented. The test

pressures ranged from 0.3 to 27 atmospheres. The isotropic confining test

pressure is in most cases equal to the estimated in-situ mean effective stress

calculated based on the sample depth, location of water table and assuming 0.5 as

the coefficient of horizontal earth pressure at rest. However, some specimens

41
were tested at more than one state of stress. As a result, the total number of test

pressures at which nonlinear measurements are performed is slightly more than

the total number of specimens. The number of samples from each geographic

region in each confining pressure category is presented in Table 3.22.

Figure 3.9 Distribution of confining pressures at which nonlinear measurements


were performed

Table 3.22 Distribution of collected according to the isotropic confining


pressure in each geographic region

Test Pressure Range (atmospheres)


Geographic Region 0.3-0.5 0.5-1.0 1.0-2.0 2.0-4.0 4.0-8.0 8.0-16.0 16.0-27.2 TOTAL
Northern California 7 6 7 12 6 2 1 41
Southern California 5 5 11 12 13 4 5 55
South Carolina 1 4 1 5 5 2 1 19
Taiwan 1 1 3 3 - - - 8
TOTAL 14 16 22 32 24 8 7 123

42
Figures 3.10 and 3.11 show the distribution of soil samples according to

the Unified Soil Classification System (USCS) designation and in terms of their

plasticity, respectively. It is important to note that soils with a wide range of

plasticity are represented in this database. About half of the soils classify as fine-

grained soils. Coarse-grained soils included in this study are limited to sands. Due

to limitations on specimen size, gravelly soils were not tested as part of this work.

The number of samples in each sample depth category divided according to their

Unified Soil Classification System (USCS) designation is presented in Table 3.23.

This table shows whether a given soil type from a given depth range is

represented in the database or not. It is important to note that a variety of soil

types from a wide range of sampling depths are represented in this database.

Figure 3.10 Distribution of soil samples according to soil type as classified by


the Unified Soil Classification System (USCS)

43
Figure 3.11 Distribution of soil samples according to soil plasticity in terms of
the plasticity index, PI

Table 3.23 Distribution of soil samples according to the Unified Soil


Classification System (USCS) designation and sample depth
Depth Range (meters)
Soil Type 0-5 5-10 10-20 20-30 30-50 50-100 100-200 200-263 TOTAL
CH 1 1 3 3 - 1 3 - 12
CL 3 5 2 1 6 6 3 2 28
CL-ML 3 - - 1 - - - - 4
MH - - - 3 - - - - 3
ML - 1 2 1 6 - 1 - 11
SC 3 1 1 1 - - - 1 7
SC-SM - 1 - 1 - 1 1 - 4
SM 3 3 7 3 4 3 1 1 25
SP 1 - 1 - - - - - 2
SP-SM - 1 1 1 3 2 1 9
SW-SC - 1 - - - - - - 1
SW-SM - 1 - - 1 2 - - 4
TOTAL 14 15 17 15 20 15 10 4 110

44
Figures 3.12 through 3.15 present distributions of soil samples according

to total unit weight, dry unit weight, water content and void ratio, respectively. In

this study, specific gravity of sandy soils is assumed 2.65 and that of clayey soils

is assumed 2.70. It is important to note that most of the soils included in this study

are competent soils with low void ratios sampled from geotechnical sites that have

not liquefied during seismic activity. Consequently, none of these samples have

exhibited major changes in their stiffness (or normalized modulus reduction

curve) due to cycling at a given strain amplitude within the range of shearing

strains testing was performed. Normally consolidated soils with higher plasticity

are the exceptions that contribute to the void-ratio diversity of this database.

Figure 3.16 shows the variation of the dry unit weights of the soil samples

with depth. In this figure, dry unit weights of the coarse grained soils (sands) are

observed to form a relatively narrow band compared to that of fine grained soils

included in this study. Also, dry unit weights of both the fine-grained and coarse-

grained soils are observed to increase with depth due to higher confining

pressures at deeper soil layers. Figures 3.17 and 3.18 present variations of water

content and void ratio with depth, respectively. These figures indicate trends

consistent with those observed in Figure 3.16. Increase in dry unit weight

necessitates a decrease in the void space (and, therefore, a decrease in water

volume filling the voids) within the soil structure. As a result, void ratio and water

content of both the fine-grained and coarse-grained soils are observed to decrease

with depth. The coarse grained soils (sands) are again observed to form a

relatively narrower band than that of the fine grained soils included in this study.

45
Figure 3.12 Distribution of soil samples according to total unit weight

Figure 3.13 Distribution of soil samples according to dry unit weight

46
Figure 3.14 Distribution of soil samples according to water content

Figure 3.15 Distribution of soil samples according to void ratio

47
3
Dry Unit Weight, gr/cm
0.00 0.50 1.00 1.50 2.00 2.50
0
50
100
Depth, m 150
200 Clayey Soils
250 Silty Soils (a)
300

3
Dry Unit Weight, gr/cm
0.00 0.50 1.00 1.50 2.00 2.50
0.0
50.0
100.0
Depth, m 150.0
200.0
250.0 Sandy Soils
(b)
300.0

Figure 3.16 Variation of dry unit weight with depth of (a) fine grained and (b)
coarse grained soils included in this study

48
Water Content, %
0 20 40 60 80 100
0
50
100
Depth, m 150
Clayey Soils
200 Silty Soils
250 (a)
300

Water Content, %
0 20 40 60 80 100
0.0
Sandy Soils
50.0
100.0
Depth, m 150.0
200.0
250.0
(b)
300.0

Figure 3.17 Variation of water content with depth of (a) fine grained and (b)
coarse grained soils included in this study

49
Void Ratio, e
0.00 0.50 1.00 1.50 2.00 2.50
0
50
100
Depth, m 150 Clayey Soils
200 Silty Soils
250 (a)
300

Void Ratio, e
0.00 0.50 1.00 1.50 2.00 2.50
0.0
Sandy Soils
50.0
100.0
Depth, m 150.0
200.0
250.0
(b)
300.0

Figure 3.18 Variation of void ratio with depth of (a) fine grained and (b) coarse
grained soils included in this study

50
In Figures 3.19 and 3.20, the distribution of samples according to

estimated overconsolidation ratio and the variation of estimated overconsolidation

ratio with depth are presented, respectively. The overconsolidated soils included

in this study are observed to be sampled from depths less than about 50 m.

Unfortunately, it was not feasible to perform consolidation tests on these samples

during the course of this work. Overconsolidation ratio of the samples are

estimated based on the characteristics of log Gmax – log σo’ relationships. The

effective isotropic confining pressure, at which a break in the log Gmax – log σo’
relationship is observed, is assumed to be the maximum mean effective stress that

the sample has experienced. Overconsolidation ratio is calculated by dividing this

pressure to the estimated mean effective stress. In most cases, the soils are

classified as normally consolidated when a clean break in the log Gmax – log σo’

relationship is not observed.

Figure 3.19 Distribution of soil samples according to estimated


overconsolidation ratio
51
Estimated Overconsolidation Ratio, OCR
0 2 4 6 8 10
0
50
100
Depth, m 150 Clayey Soils
200 Silty Soils
250
(a)
300

Estimated Overconsolidation Ratio, OCR


0 2 4 6 8 10
0.0
Sandy Soils
50.0
100.0
Depth, m 150.0
200.0
250.0 (b)
300.0

Figure 3.20 Variation of estimated overconsolidation ratio with depth of (a) fine
grained and (b) coarse grained soils included in this study

52
3.7 SUMMARY

The data that is analyzed in order to evaluate nonlinear soil behavior is

presented herein. This database has been compiled over the past decade from

research work supported by various organizations and carried out by a number of

graduate students working in the soil dynamics laboratory at The University of

Texas at Austin. Information regarding specimens from each geotechnical site in a

given geographic region is presented in separate tables showing physical

properties of soils and citing the publication that contains the original data.

A discussion regarding the distribution of samples in terms of their

geographic location, depth, soil type, plasticity index, void ratio and unit weight is

also presented at the end of this chapter in an attempt to familiarize the reader

with the characteristics of the database that is utilized in this study.

53
CHAPTER 4
OBSERVED TRENDS IN DYNAMIC SOIL PROPERTIES

4.1 INTRODUCTION

Dynamic soil properties (in terms of G and D) and the parameters that

affect these properties are discussed in this chapter. The relative importance of

each parameter on G and D and the trends reported in the literature and/or

observed during the course of this work are presented. This discussion is

presented so that the shortcomings of the existing empirical curves presented in

Chapter Five can be assessed and the strengths and limitations of the improved

empirical relationships developed in this study can be easily recognized.

4.2 BACKGROUND

Nonlinear dynamic soil properties are affected by a number of parameters

which have varying levels of importance. These parameters can be divided into

two groups: 1) parameters that relate to the static and dynamic loading conditions,

and 2) parameters that relate to the material type.

Important parameters related to the loading conditions which affect

nonlinear soil behavior are:

a) strain amplitude,

b) magnitude of the effective confinement state, often expressed by the

“equivalent” effective isotropic confining pressure,

c) duration of the effective confinement state, sometimes termed the

“long term time effect”,

54
d) number of loading cycles,

e) loading frequency (or strain rate), and

f) overconsolidation ratio (or loading history).

Soils are natural materials that can, and typically do, vary widely. The

behavior and performance of these materials tend to change significantly from

one soil to another. One of the challenges that a geotechnical engineer has to deal

with is the necessity to design an engineered structure with the material available

at a given site. As a result, for decades, geotechnical engineers have been

classifying different soils and associating their performance in various

applications with soil classes. This perspective has been utilized herein to analyze

the impact of material type on dynamic soil behavior. This study is an effort to

characterize nonlinear behavior of “competent” soils (soils that do not undergo

large volume changes during dynamic loading) at shearing strain amplitudes less

than 1 %. The results of this research are intended to be utilized in the analysis of

free-field ground motions during design level earthquakes.

Finally, the effect of sampling disturbance on dynamic soil properties is

briefly discussed in this chapter in order to show the importance of small-strain,

in-situ seismic measurements and to justify the emphasis in this work on

normalized modulus reduction curves rather than absolute values of stiffness.

Unfortunately, seismic methods have not been successfully used to date to

measure in-situ material damping ratio at any strain level on a routine basis.

Therefore, it is not possible to apply some type of laboratory-to-field

transformation to obtain field material damping curves. As a result, damping

55
curves measured in the laboratory are directly utilized as design curves in ground

motion analysis.

4.3 NONLINEAR DYNAMIC SOIL PROPERTIES

As discussed in Chapter One, soils exhibit nonlinear behavior. In other

words, secant shear modulus, G, decreases with increasing strain amplitude. Shear

modulus at small strains is referred to as small-strain shear modulus, Gmax or Go.


The relationship between shear modulus and shearing strain amplitude is typically

characterized by a normalized modulus reduction curve as shown in Figure 4.1a.

The nonlinearity in the stress-strain relationship results in an increase in

energy dissipation and therefore an increase in material damping ratio, D, with

increasing strain amplitude. Material damping ratio at small strains is referred to

as small-strain material damping ratio, Dmin. The relationship between material

damping ratio and strain amplitude is typically characterized by a material

damping curve as shown in Figure 4.1b. As noted above and as illustrated in

Figure 4.1b, D-log γ curve is expressed in absolute terms, not in normalized terms

(for instance D/Dmin-log γ or D/Dmax-log γ) because the nonlinear characteristics


of the D-log γ curve are related to the normalized modulus reduction curve of a

given soil rather than the value of material damping ratio at small or large strains.

These two curves can be broken into three strain ranges over which soils

behave differently. At small strains, γ < 0.001 %, soils exhibit linear elastic

behavior. The main source of energy dissipation is friction between particles

and/or viscosity. In other words, shear modulus is constant at a maximum value,

Gmax, and material damping ratio is constant at a minimum value, Dmin.

56
γte γtc

1.0 (a)
~0.8 N=1
G
Gmax 0.5

N=10
0
0.001 0.01 0.1 1
Shearing Strain, γ, %
γte γtc
16
(b)

D, % 8
Dmin

∆D ~3 %
0
0.001 0.01 0.1 1
Shearing Strain, γ, %

Linear Elastic
γte = elastic treshold strain
Nonlinear Elastic
γtc = cyclic treshold strain
Plastic

Figure 4.1 Linear elastic, nonlinear elastic and plastic strain ranges on (a)
normalized modulus reduction and (b) material damping curves

57
The strain amplitude at which shear modulus decreases to 98 % of its

original value is commonly called the elastic threshold strain and is denoted by
γ te . It is also called the nonlinearity threshold by Vucetic and Dobry (1991) and

Ishihara (1996). Above the elastic threshold strain, soils behave nonlinear but still

elastic. In other words, the stress-strain relationship is curved, but the

deformations are recoverable upon unloading. Due to the nonlinear stress-strain

relationship, an increase in material damping ratio is observed. The strain


amplitude at which deformations become irrecoverable is called the cyclic (or
plastic) threshold strain and is denoted by γ tc . It is also called the degradation

threshold by Vucetic and Dobry (1991) and Ishihara (1996). At this strain, shear

modulus has decreased to about 80 % of Gmax, and material damping ratio is about
3 % higher than Dmin (Stokoe et al., 1999).

Above the cyclic threshold strain, soils may change volume as they

deform. Soils exhibit different behavior when sheared depending on how dense

they are packed. Loose saturated soils tend to contract and/or develop positive

pore pressures while dense soils tend to dilate and/or develop negative pore
pressures. A change in pore pressure results in a change in effective stress and

normalized modulus reduction and material damping curves shift with each cycle

of loading and unloading, as presented in Figure 4.1. Normalized modulus

reduction curves of the soils analyzed in this study were observed to shift very

little (or not at all in most cases) with number of cycles while a considerable shift

in material damping curve was recognized. The effect of number of loading

cycles on dynamic soil behavior is discussed in more detail in Section 4.7.

58
4.4 EFFECT OF DURATION OF CONFINEMENT ON SMALL-STRAIN
DYNAMIC SOIL PROPERTIES

Figure 4.2 shows the effects of magnitude and duration of isotropic

confining pressure on the small-strain shear modulus and material damping ratio

for a typical soil specimen (UTA-1-J in Table 3.11). The variation of void ratio

with magnitude and duration of isotropic confining pressure is also presented in

this figure.

As shown in Figure 4.2a, Gmax increases as the specimen consolidates at a


given confining pressure and it also increases with increasing confining pressure.

As shown in Figure 4.2b, Dmin decreases as the specimen consolidates at a given

confining pressure and it also decreases with increasing confining pressure.

The impact of magnitude and duration of confining pressure is smaller if

the soil specimen is in an overconsolidated state (in other words, if the specimen

has been subjected to a higher confining pressure in the past) compared with the

normally consolidated state for all soils. Furthermore, these effects decrease as

overconsolidation ratio increases. If the soil specimen is normally consolidated,

magnitude and duration of confining pressure is observed to have a larger effect

on clayey soils than on sandy soils.

59
200
Isotropic Confining Pressure (a)
0.14 atm 0.27 atm
150 0.61 atm 1.22 atm
Gmax,
2.45 atm
MPa 100

50

10
(b)
8

Dmin, 6
%
4

0.64
(c)
0.62

e 0.60

0.58
Sandy Lean Clay (CL)
0.56
0 1 2 3 4
10 10 10 10 10
Duration of Confinement, t, min

Figure 4.2 Variation of (a) low-amplitude shear modulus, (b) low-amplitude


material damping ratio, and (c) void ratio with magnitude and
duration of isotropic confining pressure

60
The sandy lean clay (CL) specimen in Figure 4.2 is expected to reach 99%

consolidation in a fraction of the time period over which a change in small-strain

dynamic properties is observed. Part of the increase in Gmax and the decrease in

Dmin results from the consolidation of the specimen under the applied pressure.

However, small-strain dynamic properties continue to change after consolidation

of the specimen (in other words after the applied pressure has become the

effective confining pressure). The impact of duration of confining pressure after

primary consolidation is called the long-term time effect (or creep) and is

discussed in detail by Anderson and Stokoe (1978).

4.5 EFFECT OF EFFECTIVE CONFINING PRESSURE

4.5.1 Small-Strain (Linear) Dynamic Soil Properties

The effect of effective confining pressure, σo’, on small-strain dynamic

soil properties has been documented by various investigators (e.g., Hardin and

Drnevich, 1972a and b; Hardin, 1978; Stokoe et al., 1994; and Stokoe et al.,

1999). This effect is studied by measuring values of Gmax and Dmin (and void ratio

for that matter) after the specimen (UTA-1-J in Table 3.11) has fully consolidated

at each confining pressure. Typical results illustrating the effect of σo’ are
presented in Figure 4.3.

61
1000
(a)

σpm'
Gmax,
MPa 100

10

10
(b)

Dmin,
% 1 σpm'

0.1

0.64
σpm'
(c)
0.62

e 0.60

0.58 Sandy Lean Clay (CL)


Time = 1 day
0.56
0.1 1 10

Effective Isotropic Confining Pressure, σo , atm

Figure 4.3 Variation of (a) low-amplitude shear modulus, (b) low-amplitude


material damping ratio, and (c) void ratio with effective isotropic
confining pressure

62
Small-strain shear modulus, Gmax, increases with increasing effective

confining pressure as shown in Figure 4.3a. Overconsolidated soils tend to exhibit

some “memory” of stress history and they can be recognized from their bilinear

log Gmax – log σo’ relationships. The effective confining pressure at which a

change in the slope of log Gmax – log σo’ relationship is observed is the maximum

mean effective stress that the soil sample has ever experienced in the past and it is

indicated with σpm’ in Figure 4.3.

In the normally consolidated range, the slope of log Gmax – log σo’
relationship for most competent soils falls in a range of about 0.5 to 0.6. In the

overconsolidated range, Gmax is less sensitive to σo’, resulting in log Gmax – log

σo’ relationships with slopes smaller than 0.5.

Small-strain material damping ratio, Dmin, decreases with increasing

confining pressure as shown in Figure 4.3b. As in the case of Gmax,

overconsolidated soils tend to exhibit a bilinear log Dmin – log σo’ relationship. A

change in slope is observed at σpm’. The slope of log Dmin – log σo’ relationship in

the normally consolidated range is slightly higher than the slope in the
overconsolidated range. The variation of void ratio, e, with confining pressure is

also presented in Figure 4.3.

4.5.2 Nonlinear Dynamic Soil Properties

Over the past three decades, numerous studies have been conducted

regarding dynamic soil properties and the parameters affecting them. Various

investigators have synthesized this work and proposed nonlinear generic curves

for use in earthquake analyses (e.g., Seed et al., 1986, for sands, and Vucetic and

63
Dobry, 1991, for soils with plasticity). Most of these generic curves proposed in

previous studies were derived from dynamic measurements at effective confining

pressures around one atmosphere.

The importance of effective confining pressure on the variation of shear

modulus, normalized shear modulus and material damping ratio with shearing

strain is illustrated in Figure 4.4. These measurements were performed on a

normally consolidated silty sand (SM) specimen (UTA-1-M in Table 3.11). The

specimen was tested at the estimated in-situ mean effective confining pressure of

0.5 atm. Then, the confining pressure was increased to four times the estimated

in-situ mean effective confining pressure and the specimen was tested at 2.0 atm

again in a normally consolidated state. All results shown in Figure 4.4 were

determined using the resonant column method; hence, each measurement

involved about 1000 cycles of loading in the frequency range of 43 to 94 Hz. The

effect of number of cycles on dynamic soil behavior is discussed in Section 4.7.

Figure 4.4 illustrates that normalized modulus reduction and material

damping curves become increasingly linear as confining pressure increases. Only

a few investigations (e.g., Iwasaki et al., 1978; Kokusho, 1980; Ni, 1987; and

Ishibashi and Zhang, 1993) have considered the effect of confining pressure on

dynamic soil properties. However, most of these studies were restricted to

pressures much less than 10 atmospheres.

64
150
(a)

100
G, MPa
50

Silty Sand (SM)


0

1.2
(b)

0.8
G/Gmax
0.4

0.0

20
(c)
σo' ~ 0.5 atm
15
σo' ~ 2.0 atm

D, % 10

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 4.4 The effect of confining pressure on the variation of (a) shear
modulus, (b) normalized shear modulus, and (c) material damping
ratio with shearing strain amplitude as measured in the torsional
resonant column

65
As part of the ROSRINE project, numerous intact soil samples were

recovered over a depth range of 3 to 300 m. Some of these samples were tested

using combined resonant column and torsional shear (RCTS) equipment at

isotropic confining pressures ranging from 0.25 to 30 atmospheres. The results of

these tests show that depth as manifested through confining pressure, has a

significant impact on the shear modulus, normalized modulus reduction and

material damping curves for all soils (Stokoe et al., 1999). Typical representative

results from the ROSRINE project are illustrated in Figures 4.5 and 4.6. In

Figures 4.5a and b, average normalized modulus reduction curves for soils with

moderate plasticity and for nonplastic soils are presented, respectively. In Figure

4.6, the effect of confining pressure on normalized modulus reduction and

material damping curves is illustrated based on the results of the tests performed

on silty sands. However, only general trends were noted in the ROSRINE study.

Unfortunately, a quantitative model explaining these trends was not developed.

Hence, ROSRINE project formed the foundation for this study.

66
1.2
(a)

0.8
G/Gmax PI = 2 to 36 (%)
0.4 Depth < 7.5 m
Depth = 7.5 to 100 m
Depth = 100 to 250 m
0.0

1.2
(b)

0.8
G/Gmax
0.4 Non-Plastic Soils
Depth = 7.5 to 100 m
Depth = 100 to 250 m
0.0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 4.5 The effect of confining pressure on normalized modulus reduction


curve (a) for soils with moderate plasticity, and (b) for non-plastic
soils evaluated as part of the ROSRINE study (after Stokoe et al.,
1999)

67
1.2
(a)
0.8
G/Gmax

0.4

0.0

20
ROSRINE Study (Silty Sands)
15 σo' = 0.25 atm
σo' = 1.0 atm
D, % 10 σo' = 4.0 atm
σo' = 16 atm
5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 4.6 The effect of confining pressure on (a) normalized modulus


reduction and (b) material damping curves of silty sands evaluated as
part of the ROSRINE study (after Darendeli et al., 2001)

68
Site response analyses were carried out to evaluate the impact of modeling

confining-pressure-dependent nonlinear soil properties on predicted ground

motions (Stokoe and Santamarina, 2000 and Darendeli et al., 2001). These

analyses indicate that utilizing a family of confining-pressure-dependent curves

results in larger intensity ground motions than those predicted with average

generic curves, particularly at periods less than about 1.0 sec. In Figure 4.7, a

comparison of the predicted ground motions in terms of spectral acceleration, Sa,

for a 120-m thick silty sand deposit shaken by the Topanga motion (Maximum

Horizontal Acceleration, MHA = 0.33g) is presented. The ratio of Sa predicted


using a family of confining-pressure-dependent curves (presented in Figure 4.6)

to those predicted with average generic curves is 81 % at a period of 0.3 sec and it

is 50 % at a period of 1.0 sec. This result is more pronounced for deeper sites

subjected to higher intensity input motions due to lower damping introduced by

the pressure-dependent curves. At longer spectral periods, the response is

dominated by the overall stiffness of the site. As a result, the confining-pressure-

dependent analyses tend to predict a smaller response at longer periods due to the
more linear response modeled by these curves.

69
2.5
Pressure-Dependent 5 % Structural
2.0 Soil Properties Damping
Pressure-Independent
1.5 Soil Properties
S a, g
1.0 Input Motion

0.5

0.0
0.01 0.1 1 10
Period, T, sec

Figure 4.7 Impact on nonlinear site response of accounting for the effect of
confining pressure on dynamic soil properties (after Darendeli et al.,
2001)

4.6 EFFECT OF OVERCONSOLIDATION RATIO

Overconsolidation has an effect on the dynamic properties of soils,

particularly of those with plasticity. As an example, consider the resonant column

measurements on a kaolinite specimen presented in Figures 4.8 and 4.9. This

specimen has first been consolidated at 0.34 atm, tested at confining pressures

ranging from 0.09 to 1.36 atm in a loading sequence, and then unloaded to 0.34

atm and re-tested. These tests were performed in part to investigate the impact of

overconsolidation ratio on nonlinear behavior in both the initial loading and

unloading regions.

70
1000
Kaolinite Specimen (a)

Gmax,
100
MPa
*

10

10
(b)

Dmin,
1 *
%

* Unloading
0.1

0.80
(c)
0.78

0.76 *
e
0.74
Note: Specimen was consolidated at 0.34 atm
0.72 before testing.
0.70
0.01 0.1 1 10
Effective Isotropic Confining Pressure, σo', atm

Figure 4.8 The effect of overconsolidation ratio on the variation of (a) shear
modulus, (b) material damping ratio, and (c) void ratio with effective
isotropic confining pressure as measured in the torsional resonant
column

71
100
Kaolinite Specimen (a)
80

60
G, MPa
40

20

1.2
(b)

0.8
G/Gmax
σo' = 0.34 atm
0.4
Loading, OCR = 1.0
Unloading, OCR = 4.0
0.0

15
(c)
Shearing strains in RC test were
corrected to the average of the
10
first 3 free-vibration cycles.
D, %
5

0
-5 -4 -3 -2 -1
10 10 10 10 10
Shearing Strain, γ, %

Figure 4.9 The effect of overconsolidation ratio on the variation of (a) shear
modulus, (b) normalized shear modulus, and (c) material damping
ratio with shearing strain amplitude as measured in the torsional
resonant column

72
As illustrated in Figure 4.8 and discussed in Section 4.5, overconsolidated

soils tend to exhibit some “memory” of stress history. As a result, Gmax is larger

and Dmin is smaller in the overconsolidated state. Therefore, small-strain dynamic

properties of overconsolidated soils are less sensitive to σo’. Once a specimen is

consolidated at a higher confining pressure (back to a normally consolidated

state), a substantial decrease in void ratio results in a change in overall soil

structure and therefore a change in dynamic behavior.

The difference in Gmax for the normally consolidated and overconsolidated


states results in different nonlinear shear modulus curves as shown in Figure 4.9a.

However, in the strain range that the measurements are performed, the normalized

modulus reduction curve exhibits only a slight difference for this material as

presented in Figure 4.9b. Material damping curves for the normally consolidated

and overconsolidated states are also observed to follow a similar trend, with the

D-log γ relationship shifting slightly to higher strain amplitudes along with a

slight decrease in Dmin as shown in Figure 4.9c. Nevertheless, overconsolidation

ratio should be expected to have some impact on nonlinear soil behavior, and it
should be accounted for in developing the next generation of normalized modulus

reduction and material damping curves.

73
4.7 EFFECT OF NUMBER OF CYCLES

The effect of number of cycles on G and D can be investigated using the

combined RCTS equipment. Figure 4.10 shows a comparison of the first and tenth

cycles (N = 1 and 10) of torsional shear tests, and resonant column test results

(assuming N ~ 1000 cycles) for a typical competent soil specimen that does not

undergo large volume changes during dynamic loading (UTA-1-M in Table 3.11).

As illustrated in Figure 4.10a, shear modulus stays constant and equal to

Gmax below an elastic threshold strain, which is nominally in the range of 0.001 %
to 0.01 %. The value of Gmax measured during RC testing is slightly higher than

Gmax measured during TS testing due to the effect of frequency in this strain range

as discussed in Section 4.8. As shearing strain increases above the elastic

threshold, G decreases nonlinearly with increasing γ. Shear modulus decreases in

a similar manner in both the RC and TS tests (Stokoe et al., 1999).

Number of loading cycles, N, has no effect on G until the cyclic threshold

strain (nominally in the range of 0.01 % and 0.1 %) is exceeded. Above the cyclic

threshold strain, G varies with γ and N. The value of G somewhat decreases with
increasing N at a constant γ. The effect of N on G can be influenced by soil type,

void ratio, confining pressure, degree of saturation and soil plasticity. However,

for the “competent” soils tested in this study, N has a minor impact on G as

shown in Figure 4.10a. The variation in normalized shear modulus, G/Gmax, with

the logarithm of shearing strain is shown in Figure 4.10b. The trends, which are

related to G, can also be easily observed in the G/Gmax – log γ curves.

74
100
Silty Sand (SM) (a)
80

60
G, MPa
40

20

1.2
(b)

0.8

G/Gmax
0.4 Note:
σm' ~ 0.5 atm
0.0

20
RC (~ 1000 Cycles) (c)
15 st
TS 1 Cycle
th
D, % 10 TS 10 Cycle

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 4.10 The effect of number of loading cycles on the variation of (a) shear
modulus, (b) normalized shear modulus, and (c) material damping
ratio with shearing strain amplitude as determined in the combined
RCTS testing

75
It is seen in Figure 4.10c that material damping is constant and equal to

Dmin at strains less than or equal to the elastic threshold strain, which is nominally

equal to or slightly less than that found for G. As with Gmax, there is a difference

between Dmin values determined in the RC and TS tests because of different

loading frequencies in the two tests. (This point is discussed in the following

section.) As γ increases above the elastic threshold, D increases significantly. A

cyclic threshold strain also exists for D. The cyclic threshold for D is observed to

be somewhat smaller than that found for G. (However, this result is assumed to
show that D is more sensitive to changes in γ around γ tc than G) Above the cyclic

threshold, D decreases as N increases with the importance of N increasing with γ.

Much of the decrease in D with increasing N occurs in the first 10 cycles as

shown in Figure 4.10c. When results of resonant column tests are compared with

the data collected during the tenth cycle of torsional shear testing, the effect of N

on G and D is observed to be overwhelmed by the effect of loading frequency

(discussed in Section 4.8). It is also interesting to note that N has a greater

influence on D than G (Stokoe et al., 1994; and Stokoe et al., 1999).

4.8 EFFECT OF LOADING FREQUENCY

The effect of excitation frequency, f, on Gmax and Dmin is shown in Figure


4.11 for an intact sandy lean clay (CL) specimen (UTA-1-J in Table 3.11). In this

case, the effect of excitation frequency on Gmax is small, averaging only about 10

% as frequency increases by an order of magnitude (for frequencies ranging from

1 Hz to 100 Hz) at a given confining pressure. On the other hand, the effect of

excitation frequency on Dmin is very significant above 1Hz, with Dmin increasing

76
by about 100 % over a log-cycle increase in excitation frequency. This effect is

clearly shown in Figure 4.11b, where all values of Dmin measured in the RC test

plot above values measured in the TS test at 1 Hz. It is also important to note that

the effect of f on Dmin is more pronounced at higher frequencies.

200
Isotropic Confining Pressure (a)
0.14 atm 0.27 atm
150 TS RC
0.61 atm 1.22 atm
Gmax,
2.45 atm
MPa 100

50

10
Sandy Lean Clay (CL) (b)
8

6 TS RC
Dmin, %
4

0
0.001 0.01 0.1 1 10 100 1000
Loading Frequency, f, Hz

Figure 4.11 The effect of loading frequency on (a) low-amplitude shear modulus,
and (b) low-amplitude material damping ratio as determined in the
combined RCTS testing

77
The effect of loading frequency on Gmax and Dmin can be easily compared

when the data collected over a frequency range are normalized with the value

measured at 1 Hz. Figure 4.12 shows such a generalized summary comparison

derived from testing numerous specimens. The relative widths of the bands

indicate how much more sensitive small-strain material damping ratio is to

frequency than small-strain shear modulus.

3
Note:
Intact Specimens Increasing Plasticity
of Soils with Index, PI
PIs = 0 to 35 %
Gmax
2
Gmax 1Hz
D min /D min 1Hz
or
Inc.
Dmin PI
1
D min 1Hz
Gmax /Gmax 1Hz

0
0.01 0.1 1 10 100
Excitation Frequency, f, Hz

Figure 4.12 Comparison of the effect of loading frequency on low-amplitude


shear modulus and low-amplitude material damping ratio (from
Stokoe and Santamarina, 2000)

78
In Figure 4.13, the effect of excitation frequency on the variation of shear

modulus, normalized shear modulus and material damping ratio with shearing

strain are presented for the same clayey specimen. The G-log γ and G/Gmax-log γ

relationships are observed not to be very sensitive to frequency. On the other

hand, the effect of f on Dmin is observed to shift the D-log γ relationship over the

whole strain range that the sandy lean clay (CL) specimen (UTA-1-J in Table

3.11) is tested.

The effect of frequency is observed to be sensitive to soil type and

plasticity as discussed in the following section. Significant variability with soil

type has been observed in the effect of excitation frequency on Dmin as illustrated
by the wide band in Figure 4.12.

As a result, it may be important to capture the frequency dependence of

small-strain material damping in developing an empirical model to represent

dynamic soil behavior (particularly, for analyses that involve dynamic loading

with frequencies above 10 Hz). Unfortunately, this issue has not been addressed

in any of the generic relationships proposed in the literature.

79
100
(a)
80

60
G, MPa
40

20 Sandy Lean Clay (CL)


0

1.2
(b)

0.8

G/Gmax
0.4 Note:
σm' ~ 0.5 atm
0.0

20
RC (~ 1000 Cycles) (c)
15 st
TS 1 Cycle
th
D, % 10 TS 10 Cycle

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 4.13 The effect of loading frequency on the variation of (a) shear
modulus, (b) normalized shear modulus, and (c) material damping
ratio with shearing strain amplitude as determined in the combined
RCTS testing

80
4.9 EFFECT OF SOIL TYPE

The effect of soil type is very important when considering linear and

nonlinear dynamic soil properties as discussed in Sections 4.9.1 and 4.9.2,

respectively. This effect also manifests itself in influencing the relative effect of

other parameters such as loading frequency, number of cycles, etc. as discussed

below.

4.9.1 Small-Strain Dynamic Soil Properties

Figure 4.14 shows a comparison of the variation in Gmax and Dmin with
effective isotropic confining pressure from RCTS testing of two specimens at

similar confining pressures. One of the specimens is a silty sand (SM) specimen

(UTA-1-M in Table 3.11) and the other is a sandy lean clay (CL) specimen

(UTA-1-J in Table 3.11).

As shown in Figure 4.14a, the sandy lean clay (CL) exhibits a memory of

loading history characterized by the bilinear log Gmax – log σo’ relationship while

the silty sand (SM) follows almost a straight line. RC and TS test results denoted

with solid and open symbols in this figure are observed to be very close. This

figure also confirms that Gmax is not very sensitive to excitation frequency.
On the other hand, the values of Dmin in the log Dmin – log σo’
relationships presented in Figure 4.14b are quiet different for the two material

types. The sandy lean clay (CL) has much higher damping than the silty sand

(SM). This finding is consistent with the general trends reported in the literature

such that small-strain material damping increases with increasing soil plasticity

(Stokoe et al., 1994; and Stokoe et al., 1999).

81
1000
RC TS Soil Type (a)
Silty Sand (SM)
Sandy Lean Clay (CL)
G,
MPa 100

10

10
(b)

Dmin ,% 1

0.1
0.1 1 10
Effective Isotropic Confining Pressure, σo', atm

Figure 4.14 The effect of soil type on the variation of (a) low-amplitude shear
modulus, and (b) low-amplitude material damping ratio with
effective isotropic confining pressure as determined in the combined
RCTS testing

Unfortunately, most of the generic curves used in state of practice (e.g.,

Vucetic and Dobry, 1991) are not accurate in terms of representing this trend in

Dmin. These generic curves were synthesized from studies generally performed

with relatively older cyclic testing equipment. Because of accuracy problems at

small strains, damping measurements were not performed and data were

extrapolated to represent the small-strain behavior.

82
It is also important to note that the Dmin values measured in the RC and TS

tests are observed to be quite different for both of these materials. This is due to

the effect of excitation frequency as discussed in Section 4.8. The effect of

frequency is more pronounced on the sandy lean clay (CL). Strain-rate effects

being more pronounced in plastic soils is a well-known phenomenon reported in

the geotechnical engineering literature.

Figures 4.15 and 4.16 show the impact of excitation frequency for these

two soils in more detail. Results of torsional shear tests performed at several

frequencies well below resonance and resonant column tests performed at a

relatively higher frequency are presented for the sandy lean clay (CL) and for the

silty sand (SM) in these figures.

Figures 4.15a and 4.16a show the Gmax and Dmin measurements,
respectively, while Figures 4.15b and 4.16b present the same data using a

different perspective. The data in Figures 4.15b and 4.16b have been normalized

with the TS measurements at 1Hz in order to indicate the sensitivity of the small-

strain dynamic properties to excitation frequency.


It is important to note the scales used in Figures 4.15b and 4.16b. An

increase in Gmax on the order of 5 % to 10 % is observed per order of magnitude


increase in frequency (between 10 Hz and 100 Hz) for the silty sand (SM) and the

sandy lean clay (CL), respectively. On the other hand, an increase in Dmin on the

order of 80 % to 120 % is presented in Figure 4.16b over the same frequency

range.

83
100
(a)
80
Gmax, 60
MPa RC
40 TS

20 Silty Sand (SM)


Sandy Lean Clay (CL)
0

1.2
(b)
1.1
Gmax
Gmax 1.0 TS RC
1Hz
0.9

0.8
0.01 0.1 1 10 100
Loading Frequency, f, Hz

Figure 4.15 The effect of soil type on the variation of low-amplitude shear
modulus with loading frequency as determined in the combined
RCTS testing

In general, frequency effects on modulus may be considered small to

negligible keeping in mind: 1) variability in soil conditions and 2) how

representative a small specimen may be in estimating field stiffness. On the other

hand, a considerable change in small-strain damping, which can potentially shift

the whole D – log γ relationship, should be captured in empirical curves for

certain frequency ranges.

84
5
Silty Sand (SM) (a)
4
Sandy Lean Clay (CL)
3
Dmin ,%
2

1
TS RC
0

3
(b)

Dmin 2
Dmin
1Hz
1
TS RC

0
0.01 0.1 1 10 100
Loading Frequency, f, Hz

Figure 4.16 The effect of soil type on the variation of low-amplitude material
damping ratio with loading frequency as determined in the combined
RCTS testing

4.9.2 Nonlinear Dynamic Soil Properties

In terms of nonlinear soil behavior, the G/Gmax – log γ and D – log γ

curves for the SM and CL specimens discussed in the previous section are shown

in Figures 4.17 and 4.18. The comparison of the nonlinear behavior of the two

soils does not represent a wide range of soil types. However, Figure 4.17 indicates
a shift in normalized modulus reduction curve with changing soil type.

85
1.2

0.8
Note:
G/Gmax σm' ~ 0.5 atm
0.4
Silty Sand (SM)
Sandy Lean Clay (CL)
0.0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 4.17 The effect of soil type on the normalized modulus reduction curve as
measured in the torsional resonant column

Material damping curves determined using both the resonant column and

the torsional shear methods are shown in Figure 4.18. The data collected during

the first and the tenth cycles of torsional shear test indicates that the silty sand

(SM) is more sensitive to number of cycles than the sandy lean clay (CL). Also,

comparison of the resonant column and torsional shear data illustrates higher

sensitivity of sandy lean clay (CL) to loading frequency than the silty sand (SM).

In Figure 4.19, normalized modulus reduction and material damping

curves of five different soils with a wide range of plasticity are presented. All of

these soils were tested at similar confining pressures and both sets of curves are

observed to shift to higher strains as plasticity index, PI, increases. This trend

agrees with all empirical curves presented in the literature, which show the effect

of PI on normalized modulus reduction and material damping curves, (e.g., Sun et

al., 1988; Idriss, 1990; Vucetic and Dobry, 1991; and Ishibashi and Zhang, 1993).

86
20
Silty Sand (SM) (a)
15 Sandy Lean Clay (CL)

D, % 10

5
RC Test
0

20
st (b)
TS Test 1 Cycle
15 f = 1 Hz

D, % 10 Note:
σm' ~ 0.5 atm
5

20
th
(c)
15 TS Test 10 Cycle
f = 1 Hz
D, % 10

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 4.18 The effect of soil type on the material damping curve determined at
(a) N ~ 1000 cycles, (b) N = 1 cycle, and (c) N = 10 cycles from
combined RCTS testing

87
1.2
(a)

0.8 RC Test
G/Gmax PI = 10 %
PI = 15 %
0.4 PI = 36 %
PI = 79 %
Peat
0.0

20
σm ' ~ 0.5 atm (b)
15 RC Test

D, % 10

20
th (c)
TS Test 10 Cycle
15 f = 1 Hz

D, % 10

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 4.19 The effect of soil type on normalized modulus reduction and
material damping curves (after Stokoe et al., 1999)

88
The values of Dmin increase with increasing PI as presented in Figure 4.19.

On the other hand, the values of D at high strains (γ ~ 0.1 %) decrease as PI

increases also as shown in this figure. This rather complex relationship between

the D - log γ curves for different soils is not shown in any empirical curves. It has

been presented in a general sense in Electric Power Research Institute, EPRI

(1993b and c), based on RCTS tests of intact soil specimens tested at UT. This

behavior has also been observed by Vucetic et al. (1998) following the EPRI

study.

The general switching in the relative positions of the D- log γ curves for

the different soil types is best shown in the TS tests at an excitation frequency of 1

Hz. The effect of excitation frequency impacts the RC measurements and the

effect of number of loading cycles impacts the tenth-cycle TS measurements.

The effect of σο’ on the normalized modulus reduction and material


damping curves is discussed in Section 4.5. The general trend shows both of the

curves shifting to higher strains. In case of D - log γ relationships, the curves

simultaneously shift downward. As a result, D decreases slightly at a given γ as


σο’ increases. In general terms, the largest shift is shown by nonplastic soils, with

the effect decreasing with increasing plasticity (Stokoe et al., 1994; and Stokoe et

al., 1999). Similar behavior is observed for normalized shear modulus, except no

shift occurs at small strains with increasing σο’ because the curve is normalized

with Gmax.

89
4.10 EFFECT OF SAMPLE DISTURBANCE

Although laboratory testing methods have been standardized and

laboratory testing equipment has been significantly improved over the past several

decades, estimating accurate engineering properties of soils has always been and

will always be a challenge for geotechnical engineers. The data collected in the

laboratory should always be evaluated in terms of its success in representing the

in-situ conditions. Beside scaling effects (due to characterizing a soil deposit with

specimens that are only a few 100 cubic centimeters in volume), effects of the

sampling operation on the laboratory measurements of an engineering property

have to be taken into account prior to utilizing a laboratory test result in design.

4.10.1 Effect of Disturbance on Gmax = ρVs2

Figure 4.20 shows the shear-wave velocity measurements at one of the

sites (La Cienega located in West Hollywood) characterized as part of the

ROSRINE project. Two sets of data are presented in this figure: 1) in-situ seismic

(OYO logger and crosshole) measurements, and 2) laboratory measurements.

Comparison of the two sets of data clearly shows a discrepancy between field and
laboratory values of shear wave velocity, Vs, with laboratory Vs values generally
lower than the field measurements.

90
0

50

0
2
100
4
6
Depth, m

150 8
10
100 200 300
Vs, m/sec
200

250
OYO Logger Measurements
Crosshole Measurements
Laboratory Measurements
300
0 200 400 600 800 1000

Shear Wave Velocity, m/sec

Figure 4.20 Comparison of field and laboratory measurements of shear wave


velocity at the La Cienega site in the ROSRINE project

91
It is important to note that these samples were recovered and tested

following procedures that should be considered high-quality, state-of-the-art

practice. The ratios of laboratory measurements to field measurements at this site

range from 0.63 to 1.07. Part of the difference between the field and laboratory

values should be attributed to variability of soil conditions at the site and how

representative a small test specimen can possibly be relative to the soil deposit.

However, an important reason for the discrepancy between laboratory and

field values of Vs is due to the fact that the sampling process itself causes a

reduction in the soil stiffness by “damaging” the existing structure and

cementation of the soil material that has occurred due to aging under some state of

stress for thousands to millions of years. This phenomenon is also discussed in

Anderson and Woods (1975) and shear wave velocities measured in the

laboratory are generally characterized as being slightly less to considerably less

than the in-situ values.

A summary of 40 comparisons from the ROSRINE study is presented in

Figure 4.21. The data indicate that sampling disturbance is more pronounced in
stiffer soils. It is important to note that the shear modulus is proportional to the

square of shear wave velocity. As a result, a reduction of 40 % in shear wave

velocity due to sampling disturbance means a reduction of 64 % in small-strain

shear modulus. This comparison indicates the need for in-situ measurement of

Gmax at critical sites that are being characterized for geotechnical earthquake

engineering purposes.

92
Modulus Ratio, Gmax, lab / Gmax, field
0.10 0.25 0.50 0.80 1.00 1.50 2.00
0

Range from
ROSRINE
300 Study
VS, field, m/sec

600

General Trend
900
0.0 0.5 1.0 1.5
Velocity Ratio, VS, lab / VS, field

Figure 4.21 Variation of sampling disturbance expressed in terms of Vs, lab/Vs, field
and Gmax, lab/Gmax, field with the in-situ shear wave velocity

4.10.2 Effect of Disturbance on Dmin

Theoretically, it is possible to estimate the small-strain material damping

ratio from in-situ seismic measurements. Crosshole and downhole test results may

be used to evaluate the in-situ material damping ratio. The SASW method may be

extended to permit in-situ measurements of material damping ratio in addition to

shear wave velocity (Lai and Rix, 1998; and Rix et al., 2000). Response of

instrumented soil deposits to earthquakes and aftershocks can be analyzed in

order to estimate in-situ shear wave velocity and material damping ratio.

However, the accuracy of all of these methods is still questionable and no robust

field method exists today.

93
Most of these methods typically assume a horizontally layered system and

utilize attenuation of wave amplitude with distance from the source.

Backscattering of waves due to the contrast between soil layers and lateral

variability in the soil deposit causes significant uncertainty regarding the

estimates of in-situ material damping ratio. The quality of the estimate is further

reduced by geometric attenuation, which generally has a more significant impact

on attenuation with distance than material damping ratio.

As an example, Figure 4.22 shows independent measurements of material

damping ratio from field and laboratory tests. These field crosshole measurements

and the laboratory measurements were conducted by UT personnel (Fuhriman,

1993; and Hwang, 1997) and are presented in EPRI (1993a and b) along with a

discussion of the data collection and analysis procedures. The material damping

ratios from surface wave measurements were performed by Lai and Rix (1998)

and are generally less than those from crosshole testing possibly due to: 1)

different attenuation mechanisms which control at higher frequencies and produce

frequency-dependent damping ratios, 2) different volumes of soil sampled by the

methods, and 3) uncoupled analyses of Vs and Dmin. On the other hand, values of

damping ratio from the surface wave tests agree more closely with values from

resonant column and torsional shear laboratory tests than those estimated using

the crosshole method. These results seem to indicate that laboratory estimates of

Dmin can be used with some “judgment” for evaluation of soil deposits for

geotechnical earthquake engineering purposes. However, many more studies are

warranted in this area.

94
Damping Ratio, %
0 2 4 6 8 10
0

Depth,
m 6

12 Surface Wave
Crosshole (UT)
Resonant Column (UT)
Torsional Shear (UT)

Figure 4.22 Comparison of laboratory and field measurements of small strain


material damping ratio (from Stokoe et al., 1999)

4.10.3 Effect of Disturbance on Nonlinear Behavior

4.10.3.1 Comparison of Back-calculated Curves with Laboratory Test Results

Zeghal et al. (1995) back-calculated the variations of shear modulus and

material damping ratio with shearing strain amplitude using stress-strain histories

calculated from the free-field downhole accelerations at the Lotung site in Taiwan

from which specimens discussed in Section 3.5 were taken. The comparison of

the back-calculated nonlinear soil properties with the data collected at the

University of Texas at Austin is presented in Figure 4.23.

95
(a)

(b)
Estimates before
peak shearing strain

Statistical Fit

Figure 4.23 Comparison of nonlinear soil properties back-calculated from the


free-field downhole accelerations with the laboratory measurements
(from Zeghal et al., 1995)

96
Figure 4.23a indicates a good correlation between the normalized modulus

reduction curves estimated based on in-situ seismic response and the data

collected in the laboratory. There is a significant difference between the “field”

and laboratory material damping curves in Figure 4.23b. It is felt that this

difference can be attributed to different attenuation mechanisms and different

volumes of soil sampled by the two methods. However, if these curves were

utilized in ground motion analysis for earthquake resistant design, it is important

to note that the material damping curve measured in the laboratory would result in

a more conservative design than the field estimate.

4.10.3.2 Comparison of Test Results on Intact and Reconstituted Specimens

Disturbance was defined above as “damaging” the existing structure and

cementation of the soil material that has been aging under some state of stress for

thousands to millions of years. A similar process can be simulated in the

laboratory by breaking an “undisturbed” soil sample into small pieces, destroying

the existing soil structure completely and reconstituting it. This extreme situation

associated with disturbance is called remolding. Comparison of the measurements


on undisturbed and remolded soils can be utilized as an indicator of the sensitivity

to disturbance of a given engineering property for a given soil type.

97
As an example, consider the comparisons of dynamic soil properties

shown in Figures 4.24 and 4.25. In these figures, data collected from RCTS

testing of an undisturbed specimen and a remolded specimen are presented. The

first specimen was trimmed from a poorly graded sand (SP-SM) sample from

Idaho Falls (Stokoe et al., 1998c). Then, this specimen was remolded and a

second specimen was reconstituted from this material at a similar unit weight.

In Figure 4.24, the variations of small-strain shear modulus, small-strain

material damping ratio and void ratio are presented. The values of Gmax and Dmin

are observed to exhibit very similar relationships although this comparison

represents an extreme case of disturbance. Gmax values at lower confining

pressures are smaller for the remolded specimen since the undisturbed specimen

has some “memory” of state of stress in the field which is characterized by the

bilinear log Gmax – log σo’ relationship. Dmin values are quite similar (considering

the limited accuracy of damping measurements, they are essentially identical).

Figure 4.25 shows the comparison of the nonlinear soil behavior of the

undisturbed and remolded specimens at a mean effective confining pressure of


0.82 atm. About 20 % difference between the shear modulus of the undisturbed

and remolded specimens is easily recognized in Figure 4.25a at small-strain

amplitudes. On the other hand, the normalized modulus reduction curves

(presented in Figure 4.25b) for these two specimens are almost identical. The D –

log γ relationship (shown in Figure 4.25c) is nearly the same up to γ of about 0.01

%. At higher strains, the reconstituted specimen exhibits increasingly higher

values of D.

98
1000
(a)

Gmax,
100
MPa

10

10
(b)

Dmin, Poorly Graded Sand


% 1 with Silt (SP-SM)
3
Intact (γ t = 1.64 gr/cm )
3
Reconstituted (γ t = 1.65 gr/cm )
0.1

0.8
(c)
0.7

e 0.6

0.5

0.4
0.1 1 10

Effective Isotropic Confining Pressure, σo , atm

Figure 4.24 Comparison of the variation of (a) low-amplitude shear modulus, (b)
low-amplitude material damping ratio, and (c) void ratio with
effective isotropic confining pressure of intact (undisturbed) and
reconstituted (remolded) specimens

99
200
(a)
150
G, MPa
100

50 σo' = 0.82 atm

1.2
(b)

0.8

G/Gmax
0.4

0.0

20
Poorly Graded Sand (c)
15 with Silt (SP-SM)
3
Intact (γ t = 1.64 gr/cm )
D, % 10 3
Reconstituted (γ = 1.65 gr/cm )
t

0
-5 -4 -3 -2 -1 0
10 10 10 10 10 10
Shearing Strain, γ, %

Figure 4.25 Comparison of the variation of (a) shear modulus, (b) normalized
shear modulus, and (c) material damping ratio with shearing strain of
intact (undisturbed) and reconstituted (remolded) specimens

100
4.10.3.3 Comparison of Test Results on Companion Specimens

As part of the ROSRINE project, companion samples from a site in

Southern California (La Cienega) were tested at the University of California at

Los Angeles (UCLA) and the University of Texas at Austin (UT).

Test results in terms of the shear modulus, normalized modulus reduction

and material damping curves are presented in Figure 4.26. The double-specimen

direct simple shear, DSDSS, (Doroudian and Vucetic, 1995) tests were performed

by Prof. Vucetic and students at UCLA. These results are shown by the solid

symbols in this figure. The resonant column and torsional shear test results

performed at UT are shown by open symbols. The value of shear modulus

estimated based on the in-situ seismic crosshole measurements by UT personnel

is also presented in Figure 4.26a for comparison purposes.

Each UCLA and UT companion specimen was recovered from the same

undisturbed sample, and each companion specimen was tested at equivalent

effective stresses based on an effective coefficient of earth pressure at rest of

about 0.5. It is important to note that the RCTS and DSDSS confinement states

are isotropic and anisotropic, respectively.

101
120
RC (UT) (a)
TS (UT)
80 DSDSS (UCLA)
Crosshole
G, MPa Seismic (UT)

40

1.2
(b)

0.8

G/Gmax
0.4
Equivalent σo' ~ 1.1 atm
0.0

20
(c)
Silty Sand (SM)
15

D, % 10

0
0.0001 0.001 0.01 0.1 1 10
Shearing Strain, γ, %

Figure 4.26 Comparison of the variation of (a) shear modulus, (b) normalized
shear modulus, and (c) material damping ratio with shearing strain
measured using various equipment on companion soil samples (from
Stokoe et al., 1999)

102
The main difference between the results from the tests performed on

companion specimens exists in shear modulus. The G/Gmax – log γ relationships

are nearly identical as are the D - log γ relationships from RCTS and DSDSS

testing. The in-situ seismic value of Gmax is above values determined in the

laboratory due to sampling disturbance. In this case, the field value is about 50%

greater than the average value determined in the laboratory. Figure 4.26 shows

that normalized modulus reduction and material damping curves fall on top of

each other for the two tests with different stress states. These data seem to

indicate that as long as a soil specimen is tested at the in-situ mean effective

confining pressure, the nonlinear soil behavior (characterized by normalized

modulus reduction and material damping curves) should be determined with

reasonable precision even though the anisotropic state of stress in the field is not

duplicated. Thus, RCTS testing under isotropic confinement is a robust means of

characterizing nonlinear soil behavior.

4.10.3.4 Final Remarks on Effects of Disturbance

At this point in time, it is not possible to conduct field measurements to

evaluate shear modulus and material damping at working strains during design

level ground shaking. The data presented in this section strongly support the state-

of-practice of scaling normalized modulus reduction curve from laboratory with

in-situ Gmax. The state-of-practice of utilizing the laboratory material damping


curve without any modification obviously needs more investigation and should be

used prudently at this time.

103
However, a geotechnical engineer always has to consider the

consequences of attempting to estimate engineering properties from small

samples. Problems associated with in-situ seismic testing (backscattering of

waves due to the contrast between soil layers and lateral variability in the soil

deposit) are not errors in testing procedures. These phenomena are a part of the

dynamic response of a soil deposit too complex (and therefore too expensive) to

model in most geotechnical investigations. A discrepancy will always exist

between a model based on engineering properties measured in the laboratory and

actual field performance. Not being able to model some phenomenon does not

endorse ignoring it. Instead, the engineer has to overcome such challenges by

using judgment based on facts and experience.

4.11 SUMMARY

In this chapter, dynamic soil properties (G and D) and parameters that

affect them are discussed. Table 4.1 shows a list of these parameters and their

relative importance in terms of affecting normalized modulus reduction and

material damping curves. Trends reported in the literature and observations made

during the course of this work are summarized. The importance of accounting for

the impact of soil type and loading conditions in developing a new generation of

design curves is addressed.

104
Table 4.1 Parameters that control nonlinear soil behavior and their relative
importance in terms of affecting normalized modulus reduction and
material damping curves based on general trends observed during
the course of this study

Impact on Normalized Impact on Material


Parameter
Modulus Reduction Curve Damping Curve

Strain Amplitude *** ***


Mean Effective Confining
Pressure
*** ***
Soil Type and Plasticity *** ***
Number of Loading Cycles *+ ***++
Frequency of Loading
(above 1 Hz)
* **
Overconsolidation Ratio * *
Void Ratio * *
Degree of Saturation * *
Grain Characteristics, Size,
Shape, Gradation, Mineralogy
* *
+
*** Very Important On competent soils included in this study
++
** Important Soil Type Dependent
* Less Important

The effect of sampling disturbance on measured dynamic soil properties is

also discussed. Regarding the dynamic response of soil deposits, the data indicate

that: 1) in-situ measurement of Gmax is needed, 2) the normalized shear modulus

curves measured in the laboratory are not very sensitive to disturbance, and 3) the

material damping curves measured in the laboratory are the only estimates of

nonlinear material damping in the field at this point in time and should be used

cautiously.

105
The state-of-practice in geotechnical earthquake engineering involves

scaling a normalized modulus reduction curve from the laboratory by the in-situ

Gmax and utilizing the laboratory material damping curve as is. The findings

presented in this chapter support the adequacy of the state-of-practice provided

that the engineer accounts for discrepancies that might arise as a result of scaling

effects and variability of soil conditions at the site.

106
CHAPTER 5
EMPIRICAL RELATIONSHIPS

5.1 INTRODUCTION

Empirical curves which represent G/Gmax – log γ and D – log γ are widely

used in geotechnical earthquake engineering practice. The most common of these

curves are reviewed in this chapter. The strengths and weaknesses of the

empirical curves in estimating nonlinear soil behavior are discussed. This

discussion is presented in order to justify the need for an improved set of

empirical curves and equations that can be utilized in earthquake ground response

analyses, soil dynamics applications regarding base isolation problems, design of

machine (dynamically loaded) foundations, and in many other cases that require

prediction of strains under working loads.

5.2 HARDIN AND DRNEVICH (1972) DESIGN EQUATIONS

The first comprehensive study in which the parameters that control

nonlinear soil behavior were identified was the study by Hardin and Drnevich

(1972a and b). This study was published in the University of Kentucky reports

UKY 26-70-CE2 (Hardin and Drnevich, 1970a) and UKY 27-70-CE3 (Hardin

and Drnevich, 1970b). Table 5.1 shows the list of these parameters and their

relative importance in terms of their effect on shear modulus and material

damping based on their research.

107
Table 5.1 Parameters that control nonlinear soil behavior and their relative
importance in terms of affecting shear modulus and material
damping (Hardin and Drnevich, 1972b)

Impact on Modulus Impact on Damping


Parameter Clean Cohesive Clean Cohesive
Sands Soils Sands Soils
Strain Amplitude *** *** *** ***
Mean Effective Confining
Pressure
*** *** *** ***
Void Ratio *** *** *** ***
Number of Loading Cycles + * *** ***
Degree of Saturation * *** ** -
Overconsolidation Ratio * ** * **
Effective Strength Envelope ** ** ** **
Octahedral Shear Stress ** ** ** **
Frequency of Loading (above
0.1 Hz)
* * * **
Other Time Effects
(Thixotropy)
* ** * **
Grain Characteristics, Size,
Shape, Gradation, Mineralogy
* * * *
Soil Structure * * * *
Volume Change Due to
Shearing Strain below 0.5 %
- * - *
*** Very Important
** Less Important
* Relatively Unimportant
+ Relatively Unimportant Except for Saturated Sand
- Unknown

108
Hardin and Drnevich (1972b) also proposed that a hyperbolic relationship

can be used to relate shear stress and shearing strain in modeling dynamic soil

behavior. The Hyperbolic model, illustrated in Figure 5.1a, can be expressed as:
γ
τ= (5.1)
1 γ
+
Gmax τ max

where: τ = shear stress,

γ = shearing strain,

Gmax = small-strain shear modulus, and

τmax = shear strength of the soil.

In this model, reference strain is defined as:


τ max
γr = (5.2)
Gmax

By dividing both sides of Equation 5.1 by γ, the secant shear modulus, G, is

obtained:
1
G= (5.3)
1 γ
+
Gmax τ max

The normalized modulus reduction curve can be evaluated from Equation


5.3 by rearranging the equation as follows:
G 1
= (5.4)
Gmax γ
1+
γr

109
τ (γr,τmax)

1
Gmax

G γ
1
τ= 1 + γ
Gmax τmax
γ
a. Hyperbolic stress-strain relationship
τ
SAND

HYPERBOLIC

CLAY

γ
b. Effect of soil type on stress-strain relationship

Figure 5.1 Hyperbolic soil model proposed by Hardin and Drnevich (1972b)

110
Hardin and Drnevich (1972b) also proposed an approximate shape for the
material damping curve as:
γ
D γr
= (5.5)
Dmax γ
1+
γr

where Dmax is the maximum damping ratio of the soil that depends on soil type,
confining pressure, number of cycles and loading frequency.
Also as shown in Figure 5.1b, Hardin and Drnevich (1972b) observed that
soil type has an impact on the stress-strain relationship. Measured stress-strain
curves deviate from the simple mathematical model depending on the soil type.
As a result, they proposed to approximate observed soil behavior by distorting the
strain scale to make the measured stress-strain curve have a hyperbolic shape. For
this purpose, they defined a hyperbolic strain, γ h , which replaces the γ / γ r term

in Equations 5.4 and 5.5. Hyperbolic strain is defined as:


γ    γ  
γ h = 1 + a * exp − b *    (5.6)
γ r    γ r  

where “a” and “b” are coefficients that adjust the shape of the stress-strain curve
for soil type, number of cycles and loading frequency. Figure 5.2 shows the
normalized modulus reduction and material damping curves estimated based on
the hyperbolic model.

111
1.0 0.0

0.8 G/Gmax = 1 / (1 + γ h) 0.2


D/Dmax = 1 - G/Gmax
0.6 0.4
G/Gmax D/Dmax
0.4 0.6

0.2 0.8

0.0 1.0
0.01 0.1 1 10 100
Hyperbolic Strain, γ h

Figure 5.2 The normalized modulus reduction and material damping curves
estimated based on the hyperbolic model

The empirical equations proposed by Hardin and Drnevich (1972b)


account for the effects of plasticity index, overconsolidation ratio and confining
pressure mainly through adjusting reference strain. Effects of soil type, number of
loading cycles, loading frequency and saturation are taken into consideration by
adjusting Dmax in Equation 5.5 and the “a” and “b” coefficients in Equation 5.6.
Hardin and Drnevich (1972b) proposed graphs and equations based on their
research and experience. The complexity of the procedure in calculating the
normalized modulus reduction and material damping curves limited utilization of
this work in practice. However, their work represented an enormous step forward
in characterizing dynamic soil behavior.

112
5.3 EMPIRICAL RELATIONSHIPS

Numerous other researchers have been influenced by the Hardin and


Drnevich (1972a and b) work and have attempted to refine, improve and
generalize their results. In these other studies, “average” normalized modulus
reduction and material damping curves have been presented. Many of these
curves are widely accepted and utilized in practice. In this section, the strengths
and weaknesses of these empirical curves in estimating nonlinear soil behavior
are discussed.
The effect of mean effective confining pressure on normalized modulus
reduction curves is presented in Figure 5.3 based on resonant column and
torsional shear tests using hollow specimens (Iwasaki et al., 1978). These tests
were performed on saturated clean sand specimens under drained conditions. The
confining pressure used in RCTS testing ranged between 0.25 atm and 2.0 atm.
The results reported in Iwasaki et al. (1978) are consistent with the general trends
outlined in Chapter Four. However, the normalized modulus reduction curves are
found to be somewhat more linear compared to those observed during the course
of this study. The discrepancy is believed to result from the uniform grain size
distribution of the clean sand relative to the natural soils tested as part of this
work.

113
1.2

0.8
Iwasaki et al. (1978)
G/Gmax 0.25 atm
0.5 atm
0.4
1.0 atm
2.0 atm
0.0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 5.3 The effect of confining pressure on normalized modulus reduction


curve for Toyoura Sand (Iwasaki et al., 1978)

Results of cyclic triaxial tests on specimens made of the same saturated


clean sand are presented in Figure 5.4 (Kokusho, 1980). The effect of mean
effective confining pressure (cell pressure) was quantified for material damping
curves as well as for normalized modulus reduction curves by Kokusho (1980).
As in the case of Iwasaki et al. (1978), the range of cell pressures is restricted to
low confining stresses.
Ni (1987) also reported results of tests on clean sands (shown in Figure
5.5) using RCTS equipment at University of Texas at Austin.
The results reported in Iwasaki et al. (1978), Kokusho (1980) and Ni
(1987) are consistent with each other showing that RCTS and cyclic triaxial test
methods yield similar measurements for clean sand specimens. The normalized
modulus reduction and material damping curves are observed to shift to higher
strains and Dmin is observed to decrease with increasing confining pressure
consistent with the general trends outlined in Chapter Four.
114
1.2
(a)

0.8
G/Gmax
0.4

0.0

20
Kokusho (1980)
15 0.2 atm
0.5 atm
1.0 atm
D , % 10
2.0 atm
3.0 atm
5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 5.4 The effect of confining pressure on (a) normalized modulus


reduction, and (b) material damping curves for Toyoura Sand
(Kokusho, 1980)

115
1.2
(a)

0.8
G/Gmax
0.4

0.0

20
Ni (1987) (b)
15 0.41 atm
0.82 atm
1.63 atm
D , % 10
3.27 atm
5

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 5.5 The effect of confining pressure on (a) normalized modulus


reduction, and (b) material damping curves for non-plastic soils (Ni,
1987)

116
The sand curves, first proposed by Seed and Idriss (1970), and then re-
analyzed and re-proposed by Seed et al. (1986), are shown in Figure 5.6. These
sand curves are found to be more consistent with results measured in this study.
The upper and lower ranges in Seed et al. (1986) can be attributed to: 1)
variability in the characteristics of the granular particles (shape, size, gradation
and mineralogy), 2) variability in nonlinear soil behavior, 3) accuracy in
measurements, and 4) effect of confining pressure. The upper and lower ranges
are observed to correspond to silty sand behavior at confining pressures ranging
from about 0.25 atm to about 4 atm. The data, which Seed et al. (1986)
synthesized, are the results of tests performed on natural sands in this pressure
range.
The curves for soils with plasticity which were proposed by Sun et al.
(1988) are presented in Figure 5.7. The normalized modulus reduction curves
proposed by Sun et al. (1988) account for the effect of plasticity on nonlinear soil
behavior while the material damping curves are presented in terms of one mean
curve and a generalized range in the data over the range of soil plasticities. The
data, which Sun et al. (1988) synthesized, are also the results of tests performed at
confining pressures ranging from about 0.25 atm to about 4 atm. The lack of
correlation they found in material damping is due, at least in part, to the
difficulties that exist in performing material damping measurements, especially at
moderate to small strains.

117
1.2
(a)

0.8

G/Gmax
0.4

0.0

20
Seed et al., (1986)
Average for Sands
15
Range

D, % 10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ , %

Figure 5.6 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Seed et al. (1986)

118
1.2
(a)

0.8 Sun et al.,(1988)


PI = 5 -10 %
G/Gmax PI = 10 -20 %
0.4 PI = 20 -40 %
PI = 40 -80 %
PI = >80 %
0.0

20
Sun et al.,(1988)
15 Average for Clays
Range
D, % 10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 5.7 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Sun et al. (1988) for soils with
plasticity

119
The curves proposed by Idriss (1990) are presented in Figure 5.8. Two
“average” normalized modulus reduction curves are proposed by Idriss (1990): 1)
for sands and 2) for clays. On the other hand, a single material damping curve is
proposed for all soil types. The normalized modulus reduction curve proposed for
sands is consistent with Seed et al. (1986) while the modulus reduction curve
proposed for clays represents a high (about 50 %) plasticity clay based on the
curves proposed by Vucetic and Dobry (1991) which are discussed below. The
proposed material damping curve is similar to the lower bound curve proposed by
Seed et al. (1986). A unified material damping curve for all soil types can also be
attributed to uncertainty in damping measurements.
The curves proposed by Vucetic and Dobry (1991) are presented in Figure
5.9. The normalized modulus reduction and material damping curves proposed in
their study account for the effect of plasticity on nonlinear soil behavior.
However, the values of small-strain damping, Dmin, have been left somewhat
undefined due to the lack of small-strain data. As shown in Figure 5.9, the value
of Dmin is predicted to decrease with increasing soil plasticity, while the opposite
trend is observed during the course of this study as discussed in Chapter Four.
The data, which Vucetic and Dobry (1991) synthesized, are also the results of
tests performed at confining pressures ranging from about 0.25 atm to about 4
atm.

120
1.2
(a)

0.8

G/Gmax
0.4 Idriss (1990)
For Sands
For Clays
0.0

20
Idriss (1990)
15 For Sands and Clays

D, % 10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 5.8 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Idriss (1990)

121
1.2
(a)

0.8

G/Gmax
0.4

0.0

20
Vucetic and Dobry (1991)
15 Non-Plastic
PI = 15 %
D, % 10 PI = 30 %
PI = 50 %
PI = 100 %
5
PI = 200 %
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 5.9 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Vucetic and Dobry (1991)

122
The curves proposed by Ishibashi and Zhang (1993) are presented in
Figures 5.10 and 5.11. In their study, a set of equations, which generate
normalized modulus reduction and material damping curves changing with
confining pressure, σ o , and soil plasticity, PI, are proposed. The equations

associated with the normalized modulus reduction curve are:


G
= K (γ , PI )σ om (γ , PI )−mo (5.7)
Gmax

where:
   0.000556  0.4 
   e −0.0145 PI
1.3
m(γ , PI ) − mo = 0.272 1 − tanh ln (5.8)
   γ  

   0.000102 + n( PI )  0.492 
K (γ , PI ) = 0.51 + tanh ln   (5.9)
   γ  
 
 0.0 PI = 0
3.37 *10 −6 PI 1.404 0 < PI ≤ 15

n( PI ) =  −7 1.976 (5.10)
 7.0 *10 PI 15 < PI ≤ 70
 2.7 *10 PI
− 5 1 . 115
PI > 70

Ishibashi and Zhang (1993) proposed to associate the material damping curve
with the normalized modulus reduction curve as follows:
( )
0.333 1 + e −0.0145 PI 
1 .3
 G 
 
2
 G  
D=
2
0.586
G  − 1.547 G  + 1 (5.11)
  max   max  

123
1.2
(a)

0.8

G/Gmax
0.4

0.0

20
Ishibashi and
15 Zhang (1993)
0.25 atm
1.00 atm
D, % 10 4.00 atm
16.0 atm
5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ , %

Figure 5.10 The effect of confining pressure on (a) normalized modulus


reduction, and (b) material damping curves for non-plastic soils
(Ishibashi and Zhang, 1993)

124
1.2
(a)

0.8

G/Gmax
0.4

0.0

20
Ishibashi and
15 Zhang (1993)
Non-Plastic
PI = 50 %
D, % 10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 5.11 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Ishibashi and Zhang (1993)

125
The data, which Ishibashi and Zhang (1993) synthesized, are the results of
tests performed at confining pressures less than 10 atm. Unfortunately, the
proposed set of equations are observed to give unrealistic relationships at higher
pressures, as illustrated in Figure 5.10 by the curves at σ o = 16 atm. At high

confining pressures, these equations predict normalized shear modulus values


exceeding 1.0 at intermediate strains. At the same time, negative values of
material damping ratio may be predicted using Equation 5.11 under the same
circumstances.
The equations proposed by Ishibashi and Zhang (1993) are also observed
not to be accurate in representing the general trends related with Dmin. These
equations ignore the effect of confining pressure on Dmin, and the value of Dmin is
predicted to decrease with increasing soil plasticity, while the opposite trend is
presented in Chapter Four.
As part of a research project funded by Electric Power Research Institute,
EPRI, a total of 35 undisturbed soil samples from 5 geotechnical sites (Treasure
Island, Gilroy, Oakland Outer Harbor, San Francisco Airport and Lotung) were
tested in the soil dynamics laboratory using the combined RCTS equipment at
The University of Texas at Austin. These samples were taken from a depth range
of 3 m to 150 m and were tested over a wide range of confining pressures. The
results of these tests are also utilized in this study as discussed in Chapter Three.
The normalized modulus reduction and material damping curves based on the
EPRI (1993c) study are presented in Figures 5.12 and 5.13.

126
(a)

(b)

Figure 5.12 Variation in empirical (a) normalized modulus reduction, and (b)
material damping curves with depth (EPRI, 1993c)

127
(a)

(b)

Figure 5.13 Variation in empirical (a) normalized modulus reduction, and (b)
material damping curves with soil type (EPRI, 1993c)

128
In Figure 5.12, the shift in both the normalized modulus reduction and
material damping curves to higher strain levels and the decrease in Dmin with
increasing depth is consistent with the general trends regarding the effect of
confining pressure on nonlinear soil behavior outlined in Chapter Four.
In Figure 5.13, the effect of soil type and plasticity on the normalized
modulus reduction and material damping curves at moderate confining pressures
(at which bulk of the EPRI data is collected) is presented. As discussed in Section
4.9, an increase in Dmin and a simultaneous shift of the material damping curve to
higher strain levels with increasing PI is not shown in any empirical curves except
for the EPRI (1993c) study.

5.4 SUMMARY

The comprehensive study performed by Hardin and Drnevich (1972a and


b) is introduced at the beginning of this chapter in order to familiarize the reader
with the basis of generic curves widely used in the state-of-practice. The generic
curves presently utilized in practice are then discussed. It is important to note that
the generic curves proposed by Seed et al. (1986), Sun et al. (1988), Idriss (1990),
and Vucetic and Dobry (1991) are based on data collected at around 1-atm
confining pressure and these curves do not capture the effect of confining pressure
on nonlinear soil behavior.
Although Iwasaki et al. (1978) and Kokusho (1980) studied the impact of
confining pressure, these studies were limited to observations on clean sands
tested at low pressures.

129
The set of equations proposed by Ishibashi and Zhang (1993) account for
both soil plasticity and confining pressure on nonlinear behavior. However, these
equations are based on data collected at confining pressures less than 10 atm and
are observed to give unrealistic values at higher pressures. Also, the effect of soil
plasticity on Dmin is not represented accurately in any of the generic curves widely
used in the state-of-practice.
The empirical curves from the EPRI (1993c) study are based on data
collected over a relatively wider range of confining pressures and are consistent
with the general trends outlined in Chapter Four. Although the EPRI (1993c)
study is one of the most comprehensive studies of nonlinear soil behavior, the
effects of some of the factors such as loading frequency and number of cycles are
not accounted for as part of this work.
The new empirical curves based on the four-parameter soil model
discussed in Chapter Six are formulated to accurately represent (consistent with
the general trends outlined in Chapter Four) the effects of soil type, confining
pressure, loading frequency and number of cycles on the normalized modulus
reduction and material damping curves. These factors are shown to be the key
factors affecting nonlinear dynamic soil behavior for “competent” soils as
summarized in Table 4.1.

130
CHAPTER 6
PROPOSED SOIL MODEL

6.1 INTRODUCTION

In this chapter, a four-parameter model that can be used to characterize


normalized modulus reduction and material damping curves is proposed. This
model is used to develop empirical curves that are based on the hyperbolic soil
model originally developed by Hardin and Drnevich (1972b). The basic
hyperbolic relationship between stress and strain is slightly modified in order to
accommodate a better fit to the modulus reduction curves measured in the
laboratory.
The equation for the material damping curve is related to the shape of the
modulus reduction curve assuming the validity of Masing behavior (Masing,
1926) combined with two modifying parameters. To start, Masing behavior is
used to calculate material damping by evaluating the hysteresis loops that should
form for a given modulus reduction curve and two-way stress reversals. This
material damping curve is then modified using two parameters to fit the
laboratory data as discussed below.
A parametric study is also presented in this chapter to assist the reader in
becoming familiar with the modified hyperbolic model.

131
6.2 NORMALIZED MODULUS REDUCTION CURVE

The hyperbolic model proposed by Hardin and Drnevich (1972b) and the
modification of the model is discussed in this section. A modified hyperbolic
model is utilized to evaluate and model dynamic soil properties in this study.
As discussed in Chapter Five, a normalized modulus reduction curve
based on the hyperbolic model can be expressed as:
G 1
= (6.1)
Gmax γ
1+
γr

It is easy to see that reference strain, γr, corresponds to the strain amplitude when
shear modulus reduces to one half of Gmax. If one uses this approach in defining
γr, reference strain for any given normalized modulus reduction curve can be
easily evaluated from laboratory measurements as long as G/Gmax values around
0.5 are measured during testing. In fact, γr = γG/Gmax=0.5 is a key characteristic of
the hyperbolic model as employed in this research.
In this study, a relatively simple approach is utilized to fit measured stress-
strain curves. A curvature coefficient, a, is integrated into the normalized modulus
reduction curve (Darendeli, 1997) as follows:
G 1
= a
(6.2)
Gmax γ 
1 +  
γ r 

The curvature coefficient, as the name implies, has an impact on the curvature of
the normalized modulus reduction curve. The reference strain still corresponds to
the strain amplitude when shear modulus reduces to one half of Gmax. The
advantage of this modification is its simplicity. However, depending on the value

132
of the curvature coefficient, the calculated stress-strain curve may not be
asymptotic to the horizontal line defined by τmax. Since this study is an effort to
model strain amplitudes far below failure, the success of this equation in
modeling soil behavior when full shear strength is mobilized is of lesser concern.
Figure 6.1 shows a normalized modulus reduction curve calculated using
this modification to the Hardin and Drnevich (1972b) hyperbolic model; that is
using Equation 6.2 with γr = γG/Gmax=0.5 = 0.03 % and a = 0.90.
The modified hyperbolic relationship can be used to approximate the
normalized modulus reduction curve of all competent soil types at strains even in
excess of 0.3 % as shown in Chapter Eight.

1.0

G/Gmax 0.5

γ r = γ G/G = 0.03 %
max =0.5
a = 0.90
0.0
0.0001 0.001 0.01 γr 0.1 1
Shearing Strain, γ, %

Figure 6.1 Normalized modulus reduction curve (of a silty sand at 1 atm
effective confining pressure) represented using a modified
hyperbolic model

133
6.3 NONLINEAR MATERIAL DAMPING CURVE

Material damping can be assumed to result from at least two separate


phenomena. Part of the energy applied to a soil body is attenuated due to the
friction and/or viscous losses at the contact surfaces between particles. In other
words, regardless of the strain amplitude, some energy loss and, therefore,
equivalent damping should be anticipated. On the other hand, soils are extremely
nonlinear in the strain range of interest (0.0001 % < γ < 1 %) during a design-
level earthquake. The nonlinearity in the stress-strain relationship results in
energy loss in a system. Thus, a second component has to be considered. In this
study, these two sources of damping are handled separately and added to each
other in order to evaluate equivalent damping of soils.
Since material damping resulting from nonlinearity of the stress-strain
relationship is the major component, it is discussed first in this section. As
discussed in Section 6.2, the stress-strain and modulus reduction curves are
directly related to each other for two-way cyclic loading (complete stress
reversals). Once a soil is characterized in terms of its normalized modulus
reduction curve and small-strain shear modulus, it is possible to predict the stress-
strain path that the soil is expected to follow under monotonic loading. Figure
6.2a shows the normalized modulus reduction curve (Figure 6.2a is the same
curve as the one shown in Figure 6.1). This curve is scaled to Gmax = 45 MPa in
Figure 6.2b in order to evaluate the G – log γ relationship. Shear stress can be
expressed as:
τ = G *γ (6.3)

134
1.2
(a)

0.8
G/Gmax
0.4 γ r = 0.03 %
a = 0.90

0.0

50
(b)
40

30
G, MPa
20

10

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

0.015

0.010
τ,
MPa
0.005

(c)
0.000
0.0 0.2 0.4 0.6 0.8 1.0
Shearing Strain, γ, %

Figure 6.2 Stress-strain curve (of a silty sand at 1 atm effective confining
pressure) estimated based on a modified reference strain model

135
and the stress-strain curve is easily calculated from the G – log γ relationship. The
resulting stress-strain curve is presented in Figure 6.2c. This curve is theoretically
the stress-strain path under monotonic loading for the material characterized by
the normalized modulus reduction curve in Figure 6.2a and Gmax = 45 MPa.
Masing (1926) assumed that the stress-strain path during cyclic loading
could be related to the monotonic loading stress-strain path, which is also called
the backbone curve. His first attempt to relate these two stress-strain paths is
typically called “Masing behavior”. As presented in Figure 6.3, Masing behavior
assumes that hysteresis loops for two-way cyclic loading can be constructed by
scaling the backbone curve by a factor of two. After initial loading, the scaled
curve is flipped on the horizontal and vertical axes, respectively, and placed at the
end of the backbone curve to represent the unloading path. In order to represent
reloading, the scaled curve is placed at the end of the unloading path. As
unloading and reloading is continued, the same stress-strain path is followed. This
simple approach is powerful in explaining the mechanism that causes energy loss
and the formation of hysteresis loops. However, its direct application in modeling
nonlinear soil behavior is limited and has been shown to perform poorly in
various strain ranges as discussed below.

136
0.015

0.010

0.005

τ, 0.000
MPa

-0.005

-0.010

-0.015
-0.3 -0.2 -0.1 0.0 0.1 0.2 0.3

Shearing Strain, γ, %

Figure 6.3 Hysteresis loop estimated by modeling stress-strain reversals for


two-way cyclic loading according to Masing behavior

One of the major shortcomings of the Masing behavior assumption is due


to the fact that the forcing function during an earthquake is not sinusoidal and a
few more rules have to be adapted in order to utilize this approach in modeling
loadings that do not cycle between two equal and opposite values of stress (two-
way cyclic loading). Since this study concentrates on relating the material
damping curve to modulus reduction curve, the assumption of cyclic loading is
appropriate and an irregular forcing function is not relevant to this work. An
algebraic expression can be written assuming sinusoidal loading conditions to
evaluate equivalent viscous damping as a function of strain amplitude.

137
As discussed in Section 2.4 and presented in Figure 6.4, damping ratio can
be calculated using the ratio of the dissipated energy to stored strain energy in one
complete cycle of motion. Assuming Masing behavior, the area inside the
hysteresis loop (AL) can be calculated by integrating the stress-strain curve over
one loading cycle.

Shear G
Stress, τ
1

AT

AL Shearing Strain, γ

G=τ/γ
D = AL / (4 π AT)

Figure 6.4 Calculation of damping ratio utilizing a hysteresis loop

With Equations 6.2 and 6.3, the stress-strain curve can be expressed as:
γ
τ= a
* Gmax (6.4)
γ 
1 +  
γ r 

The area inside the hysteresis loop can be related to the integral of the stress-strain
curve as follows:
 1 
AL = 8 *  ∫ τdγ − τγ  (6.5)
 2 

138
Equivalent viscous damping is expressed as:
AL
Deq = (6.6)
4ΠAT

where: Π = pi (= 3.1416),
AT = stored strain energy (τγ/2),
AL = dissipated energy, and
Deq = equivalent viscous damping.
By combining Equations 6.5 and 6.6, Masing-behavior damping can be written as
a function of strain amplitude as follows:
 1 
8 *  ∫ τdγ − τγ 
 2 
DMa sin g = (6.7)
1
4Π τγ
2
By substituting Equation 6.4 into Equation in 6.7, Masing-behavior
damping can be rewritten as follows:
γ 1 γ2
∫   a dγ − 2  
a
γ γ
1 +   1 +  
DMa sin g
4
= * γ r  γ r  (6.8)
Π γ2
a
γ 
1 +  
γ r 

For a curvature coefficient, “a”, equal to 1.0, Equation 6.8 reduces to:
  γ + γ r  1 γ 2
γ r γ − γ r ln  −
  γ r  2 1 + γ
4 γr
DMa sin g ,a =1.0 = * (6.9)
Π γ2
γ
1+
γr

139
and further rearrangement of this equation results in:
 γ +γ r  
 γ − γ r ln  
100   γr  
DMa sin g ,a =1.0 (%) = 4 −2 (6.10)
Π  γ2 
 
 γ +γr 

Unfortunately, the integration in Equation 6.8 can not be evaluated


algebraically for most curvature coefficient values other than 1.0. As a result, a
numerical approach is utilized to calculate functions that represent Masing
damping for different values of the curvature coefficient.
First, damping based on Masing behavior for any given strain amplitude is
assumed to be a function of: 1) Masing damping at that strain for a curvature
coefficient equal to 1.0, and 2) the value of curvature coefficient. Thus, the
expression below is assumed to be valid:
DMa sin g = c1 DMa sin g ,a =1.0 + c 2 DMa sin g ,a =1.0 2 + c3 DMa sin g ,a =1.0 3 (6.11)

In other words, Masing damping is assumed to be a polynomial function of


Masing damping for curvature coefficient equal to 1.0, and c1, c2 and c3 are
assumed to be functions of curvature coefficient.
Masing damping for curvature coefficients ranging from 0.7 to 1.3 are
calculated through numerical integration using the trapezoid rule. These damping
curves are fitted using the expression in Equation 6.11. Thus, c1, c2 and c3 values
are calculated for sixty damping curves that are calculated based on different
values of the curvature coefficient. In Figure 6.5, variations of c1, c2 and c3 with
the curvature coefficient are presented. The solid curves passing through the data
points are the best-fit polynomial relationships.

140
1.1
(a)
1.0

c1 0.9
2
c1 = -1.1143a + 1.8618a + 0.2523
0.8 2
R = 0.9997

0.7

0.04
2
c2 = 0.0805a - 0.0710a - 0.0095
2
0.02 R = 1.0000
c2

0.00
(b)
-0.02

0.0002
(c)
0.0001

0.0000
c3
-0.0001 2
c3 = -0.0005a + 0.0002a + 0.0003
-0.0002 2
R = 0.9996
-0.0003
0.7 0.8 0.9 1.0 1.1 1.2 1.3
Curvature Coefficient, a

Figure 6.5 Variations of c1, c2 and c3 with curvature coefficient, a

141
In this way, a simple algebraic expression for a damping curve based on a
modified hyperbolic stress-strain curve and Masing behavior was derived. This
result is expressed as:

DMa sin g = c1 DMa sin g ,a =1.0 + c 2 DMa sin g ,a =1.0 2 + c3 DMa sin g ,a =1.0 3

where:
 γ +γ r  
 γ − γ r ln  
DMa sin g ,a =1.0 (%) =
100 
4  γr  − 2 (6.12)
Π  γ2 
 
 γ +γr 

c1 = -1.1143a 2 + 1.8618a + 0.2523


c 2 = 0.0805a 2 - 0.0710a - 0.0095
c3 = - 0.0005a 2 + 0.0002a + 0.0003

Figure 6.6 shows a damping curve estimated using Equation 6.12. It is


important to note that this curve indicates larger damping ratios at high strains
than experimental results observed in the course of this study and values reported
in the literature (e.g., Seed et al., 1986; Vucetic and Dobry, 1991; etc.). Also,
damping calculated based on Masing Behavior lacks small-strain damping, Dmin.
As a result, further work is required to fit the mathematical expressions to the
experimental observations. The corrections on the damping curve calculated
based on Masing behavior can be summarized as follows:
D = F * DMa sin g + Dmin (6.13)

where F is a function that adjusts damping at high strains.

142
60
Damping values calculated
through numerical intagration
40 Damping values calculated
D, % using polynomial approximation
20
γ r = 0.03 %
a = 0.90
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 6.6 Damping curve estimated based on Masing behavior

Stokoe and Lodde (1978) have shown that small-strain properties are
affected by prior high-amplitude cycling. Gmax tends to decrease slightly and Dmin
tends to increase slightly after high-amplitude cycling as shown in Figure 6.7. The
decrease in Gmax should be expected to result in a considerable decrease in the
nonlinear component of material damping. Also, on various geotechnical
earthquake engineering applications (e.g., Stokoe et al., 1998a), a cap on high-
amplitude damping is required in seismic evaluations of critical structures (in
order to be conservative). As a result, function F in Equation 6.13 is replaced with
a function of G/Gmax in the form of:
p
 G 
F = b *   (6.14)
 Gmax 

where b and p are parameters that control the characteristics of this function.

143
Figure 6.7 Effect of high-amplitude cycling on low-amplitude shear modulus
and material damping ratio (from Stokoe and Lodde, 1978)

Figure 6.8 shows the variation of this function with shearing strain amplitude for
different values of “p”. Using the value of 0.1 for the “p” parameter was observed
to fit the experimental data. In order to simplify the model, the “p” parameter was
replaced with this constant value. Thus, damping adjusted using function F is
expressed as:
0.1
 G 
D Adjusted = b *   * DMa sin g (6.15)
 Gmax 

where: b = scaling coefficient,


G/Gmax = normalized modulus, and
DAdjusted = scaled and capped material damping.

144
1.0
p
F = b * (G/Gmax)
0.8
b = 0.56
0.6
F
0.4 p=0
p = 0.1
0.2 p = 0.25
0.0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 6.8 Comparison of the variation in F with shearing strain for different
values of p

This function acts as a damping cap at very high strains. At the same time,
within this function, a parameter, which is called the scaling coefficient, is
introduced to the model. The scaling coefficient is, in a sense, the ratio of the
measured damping to the damping value which is estimated from Masing
behavior at intermediate strain amplitudes. Hence, the adjusted damping curve in
Figure 6.9b is estimated.
As discussed earlier in this chapter, small-strain damping is also accounted
for in this model. Dmin is added to the adjusted damping curve and the whole
curve is shifted by this amount as shown in Figure 6.9c. The effect of high-
amplitude cycling on Dmin is ignored in this model due to two reasons: 1) for
design purposes, it is always conservative to underestimate material damping
ratio, and 2) even a 50 % increase in Dmin (as shown in Figure 6.7 at γ = 0.1 %)
has a negligible impact on the material damping values at strain levels produced
during design level shaking.

145
60
γ r = 0.03 %
40 a = 0.90
D, %

20

(a)
0

25
0.1
20 F = b * (G/Gmax)
b = 0.56
15
D, %
10

5
(b)
0

25
Masing Behavior
20 Adjusted Curve
15 Shifted Curve
D, %
10
Dmin = 0.90 %
5
(c)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 6.9 (a) Damping curve estimated based on Masing behavior, (b) adjusted
curve using the scaling coefficient, and (c) shifted curve using the
small-strain material damping ratio

146
6.4 PARAMETRIC STUDY OF THE SOIL MODEL

In this section, a parametric study of the soil model is presented so that the
effect of each parameter on the normalized modulus reduction curve and the
material damping curve can be easily visualized. The ability of this four-
parameter model in representing the trends discussed in Chapter Four is also
discussed herein.

6.4.1 Reference Strain, γr

The effect of reference strain on the normalized modulus reduction, stress-


strain and material damping curves is presented in Figure 6.10. As discussed in
Sections 4.5 and 4.9, increasing confining pressure and increasing soil plasticity
both cause shifts to higher strain levels in the normalized modulus reduction and
material damping curves. This trend is accounted for by adjusting this parameter
as shown in Figure 6.10.

6.4.2 Curvature Coefficient, a

The effect of the curvature coefficient on the normalized modulus


reduction, stress-strain and material damping curves is presented in Figures 6.11
through 6.13. Hardin and Drnevich (1972b) have pointed out a relationship
between the shape of the stress-strain curve and soil type, which can be accounted
for with this parameter as shown in these figures.

147
1.0
(a)
0.8
0.6 Reference Strain, γ r
G/Gmax
0.4 1.00E-2 %
3.00E-2 %
0.2 1.00E-1 %
0.0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

0.03
(b)

τ, 0.02
MPa
0.01

0.00
0.0 0.2 0.4 0.6 0.8 1.0
Shearing Strain, γ, %

30

20
D, %
10
(c)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 6.10 Effect of reference strain on (a) normalized modulus reduction, (b)
stress-strain, and (c) material damping curves

148
1.0

0.8

0.6
G/Gmax
0.4 Curvature Coefficient, a
0.8
0.2 1.0
1.2
0.0 γr
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 6.11 Effect of the curvature coefficient on the normalized modulus


reduction curve

0.008 0.03

0.006
0.02
τ, τ,
0.004 MPa
MPa
0.01
0.002
(a) (b)
0.000 0.00
0.00 0.01 0.02 0.03 0.04 0.0 0.2 0.4 0.6 0.8 1.0
Shearing Strain, γ, % Shearing Strain, γ, %

Figure 6.12 Effect of the curvature coefficient on the stress-strain curve (a) at
small and intermediate strains, and (b) at high strains

149
30

20

D, %
10

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 6.13 Effect of the curvature coefficient on the material damping curve

6.4.3 Small-Strain Material Damping Ratio, Dmin

The effect of Dmin on the material damping curve is presented in Figure


6.14. As discussed in Sections 4.5, 4.8 and 4.9, confining pressure, loading
frequency and soil plasticity affect the small-strain material damping ratio. The
impact of all of these parameters on the material damping curve can be accounted
for with Dmin, which results in a general shifting of the material damping curve.
The largest relative impact of this parameter on the value of D is, of course, at
small strains.

150
30
Small Strain Damping, Dmin
0.0 %
20
0.9 %
D, % 3.0 %
10

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 6.14 Effect of Dmin on the material damping curve

6.4.4 Scaling Coefficient, b

The effect of the scaling coefficient on the material damping curve is


presented in Figure 6.15. As discussed in Section 4.7, the number of loading
cycles affects the damping curve, which can be accounted for with this parameter
by varying b with number of cycles.

151
30
Scaling Coefficient, b
0.40
20 0.56
0.70
D, %
10

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 6.15 The effect of scaling coefficient on material damping curve

6.5 SUMMARY

In this chapter, a set of equations based on the Hardin and Drnevich


(1972b) hyperbolic model is proposed to represent the general trends of dynamic
soil properties outlined in Chapter Four.
This empirical model utilizes a modified hyperbolic soil model in order to
represent the stress-strain relationships and, therefore, the normalized modulus
reduction curves. The modified hyperbolic soil model utilizes two parameters
(reference strain and curvature coefficient) and can be expressed as:
G 1
= a
(6.16)
Gmax γ 
1 +  
γ r 

152
In case of the material damping curve, the normalized modulus reduction
curve and Masing behavior are employed as a starting point. A material damping
curve is calculated by evaluating the hysteresis loops that form for a given
modulus reduction curve assuming Masing behavior. A function, F, is defined as:
p
 G 
F = b *   (6.17)
 Gmax 

which utilizes the normalized modulus reduction curve and two parameters.
Parameter “p” is replaced by a constant value of 0.1 to simplify the model. Thus,
the damping curve estimated from Masing behavior is scaled to fit the
experimental observations using the normalized modulus reduction curve and a
parameter called the scaling coefficient. The adjusted (scaled) damping curve is
then shifted by a second parameter, Dmin. This way, using two additional
parameters, the material damping curve is expressed as:
0.1
 G 
D = b *   * DMa sin g + Dmin (6.18)
 Gmax 

Hence, a four-parameter model (γr, a, b and Dmin) is used to characterize


the normalized modulus reduction and material damping curves.
A parametric study is also presented to assist the reader in becoming
familiar with the modified hyperbolic model and the impact of each of the
parameters. The general ability and usefulness of this four-parameter model in
representing the trends discussed in Chapter Four is also discussed.

153
CHAPTER 7
STATISTICAL ANALYSIS OF COLLECTED DATA USING
FIRST-ORDER, SECOND-MOMENT BAYESIAN METHOD

7.1 INTRODUCTION

In this chapter, the statistical analysis carried out as part of this study is
presented. The statistical tool utilized in this study is briefly explained and the
advantages of using this approach in geotechnical engineering are discussed.
Bayes’ theorem is introduced early in the chapter in order to familiarize
the reader with the concept of systematically utilizing both experience and
observations in statistical analysis. Then, application of the theorem to discrete
and continuous problems is briefly discussed. A discrete example is presented in
order to clarify the methodology. The computational drawbacks in direct
application of Bayes’ theorem are pointed out and an iterative method, which is
used in analyzing the collected data, is presented.
Each parameter of the soil model (presented in Chapter Six) is expressed
in the form of an equation. The equations are formulated so that the impact of
strain amplitude, effective isotropic confining pressure, loading history and
loading frequency on dynamic soil behavior can be properly modeled following
the general trends discussed in Chapter Four. These equations and the form of the
anticipated covariance structure, which describes the deviations between the data
and the model in terms of their magnitude and interrelationships (correlations),
are also presented herein.

154
7.2 BAYESIAN APPROACH

Whenever an engineer faces a problem, he or she attempts to analytically


structure the problem and represent the physical phenomenon by a mathematical
relationship. Once the parameters of the mathematical relationship are
determined, calculations are performed. The solution of the problem is achieved
when the numerical result is expressed as a physical quantity. This method is the
way a structural engineer decides the required reinforcement for a concrete beam
and it is the same method a geotechnical engineer follows to determine the
required pile length for a deep foundation. If the parameters of the mathematical
relationship are estimated accurately, the outcome is a success. The engineer can
develop a safe and economic solution.
In statistical analysis, accurate estimates of parameters require large
amounts of data. In most engineering applications, observed data are limited due
to the fact that data are an expensive commodity. As a result, using judgment and
intuition has always been an essential part of decision making in engineering.
Complex engineering problems require combining information based on
experience with observational data. The Bayesian method is a systematic way of
achieving this goal.
By modeling the unknown parameters as random variables, uncertainty
associated with the estimation of the parameters can be combined with the
inherent variability of the basic random variable through Bayes’ theorem. Thus,
engineering judgment and experience can be incorporated with observed data to
obtain a balanced estimation (Ang and Tang, 1975 and 1990).

155
7.2.1 Bayes’ Theorem

The numerical measure of the likelihood of occurrence of an event, Ei,


relative to all possible outcomes is called the probability of this event, and it is
denoted as P(Ei).
If the probability of an event depends on the occurrence of another event,
the associated probability is called conditional probability. The conditional
probability of E1 assuming E2 has occurred is denoted as P(E1/E2) and it is
expressed as:
P( E1 E 2 )
P ( E1 / E 2 ) = (7.1)
P( E 2 )

where P(E1E2) denotes probability of occurrence of both events.


P(E1E2) can also be expressed in terms of the conditional probability of E2
assuming E1 has occurred using the same form presented in Equation 7.1 as
follows:
P( E1 E 2 )
P ( E 2 / E1 ) = (7.2)
P( E1 )

By substituting Equation 7.2 into Equation 7.1, the following equation can
be written:
P( E 2 / E1 ) * P( E1 )
P ( E1 / E 2 ) = (7.3)
P( E 2 )

Considering n mutually exclusive and collectively exhaustive events, E1,


E2, …. , En; and another event A in the same sample space, we can write:
n
P ( A) = ∑ P( A) * P( Ei ) (7.4)
i =1

156
Using the definition of conditional probability, Equation 7.4 can be
expressed as follows.
n
P ( A) = ∑ P( A / Ei ) * P( Ei ) (7.5)
i =1

Therefore, if the event A has occurred, the probability of occurrence of a


particular event Ei can be expressed by Equation 7.6, which is also known as
Bayes’ theorem, and is expressed as:
P( A / Ei ) * P( Ei )
P ( Ei / A) = n (7.6)
∑ P( A / E j ) * P( E j )
j =1

7.2.2 Discrete Case

Application of Bayes’ theorem for a discrete problem in order to update an


initial estimate in light of new observations is described in this section.
If a variable X can only take on certain discrete values of x, it is called a
discrete random variable. The outcome of rolling a dice represents such a
variable. It is possible to roll only integer values between 1 and 6. In this case,
the function expressing P(X = x) for all x is utilized to describe the probability
distribution of X. This function is called a probability mass function (PMF).
If a parameter X is often used by an engineer, he or she would have some
prior knowledge of possible values and values more likely to occur than others
based on experience or intuition. This initial information can be represented in
terms of a PMF if X is a discrete random variable. Once additional information
becomes available, the prior assumptions on the parameter are modified or

157
updated for the specific problem at hand. This task can be systematically achieved
by employing Bayes’ theorem using:
P(ε / X = xi ) * P( X = xi )
P ( X = xi / ε ) = n (7.7)
∑ P(ε / X = xi ) * P( X = xi )
i =1

where P(ε / X = xi ) = the conditional probability of obtaining a

particular experimental outcome, ε, assuming


that parameter X is equal to xi,
P ( X = xi ) = the probability of parameter X being equal to xi

before the availability of experimental


information, ε, and
P( X = xi / ε ) = the probability of parameter X being equal to xi

after the initial assumptions are revised in the


light of the experimental outcome, ε.
If the prior and posterior probabilities are denoted as P’(X = xi) and P’’(X
= xi), respectively, Equation 7.7 is expressed as follows:
P(ε / X = xi ) * P' ( X = xi )
P ' ' ( X = xi ) = n (7.8)
∑ P(ε / X = x j ) * P' ( X = x j )
j =1

This equation represents the posterior PMF in light of the experimental outcome.
In order to clarify Bayesian approach, a simple example is presented
below:

To determine the insurance rate for a five-story reinforced concrete


residential building in Istanbul, Turkey, an insurance company is
in need of an estimate of seismic risk of the structure during a
design level ground motion. Unfortunately, available information
is very limited. There are two groups of reinforced concrete

158
buildings in Istanbul. Group 1 represents 30 percent of the
buildings and they do not meet the local building design code.
Forty percent of these buildings are expected to collapse during the
design level shaking. The remaining buildings meet the code and
only 1% collapse is expected for this second group. Survival of a
building during any given earthquake does not necessitate good
performance during the following one. Given the fact that the
building of interest has survived two earthquakes in 1999
(resulting in design level shaking at the site), what is the
probability of collapse for this structure during the next
earthquake?
Once the problem is structured analytically, the building of interest is
expected to fall into one of the two categories. The category of the building is a
discrete variable and the probability distribution for the two categories is the prior
PMF as shown in Figure 7.1. If the building is a member of Category 1, it is more
likely to collapse during design level shaking as tabulated in Table 7.1.

100%

70%
75%
Probability

50%
30%
25%

0%
Category 1 Category 2
Building Category

Figure 7.1 Prior probability mass function for the discrete example

159
Table 7.1 Prior information provided in the discrete example

P’(Category = 1) 30 %
P’(Category = 2) 70 %
P(collapse / Category = 1) 40 %
P(collapse / Category = 2) 1%

Based on this information, the probability of collapse for a particular


building (not knowing whether or not it was built following the code) can be
calculated as follows:
p’(collapse) = 0.30 * 0.40 + 0.70 * 0.01 (7.9a)
p’(collapse) = 12.7 % (7.9b)
The statement regarding survival of the building during two earthquakes in 1999
is observational data. Assuming survival of the building at two instances are
statistically independent:
P(success) = 1 – P(collapse) (7.10)
P(success 2 consecutive earthquakes) = P(success)2 (7.11)
Based on Equation 7.11, the prior PMF can be updated using Equation 7.12 based
on Bayes’ theorem as:
P(ε / Category = 1) * P' (Category = 1)
P' ' (Category = 1) = 2
(7.12a)
∑ P(ε / Category = i) * P' (Category = i)
i =1
P(ε / Category = 2) * P' (Category = 2)
P' ' (Category = 2) = 2
(7.12b)
∑ P(ε / Category = i) * P' (Category = i)
i =1

160
Figure 7.2 illustrates the posterior (updated) PMF based on independent
observations regarding performance of the structure during two earthquakes in
1999. Using the posterior PMF, the probability of collapse for a particular
building can be recalculated as:
p’’(collapse) = 0.136 * 0.40 + 0.864 * 0.01 (7.13a)
p’’(collapse) = 6.3 % (7.13b)

100%
86.40%

75%
Probability

50%

25% 13.60%

0%
Category 1 Category 2
Building Category

Figure 7.2 Posterior probability mass function for the discrete example

In the light of the observational data, the likelihood of the building of


interest being a member of Category 2 has increased and, as a result, the
probability of collapse updated based on this data (6.3 %) is smaller than the
estimate based on prior information (12.7 %).

161
7.2.3 Continuous Case

In most cases, the variable of interest does not take on certain discrete
values. Instead, it can take on any value within a continuum. These kinds of
parameters are called continuous random variables. The unit weight of a soil
represents such a variable. In this case, probabilities are associated with intervals.
In other words, the probability of X taking a value between x1 and x2 is relevant.
Thus, a probability density function (PDF) is defined for continuous random
variables. If fx(x) is the PDF of X, the probability of X taking a value between x1
and x2 is:
x2
P ( x1 < X ≤ x 2 ) = ∫ f x ( x)dx (7.14)
x1

In other words,
P ( x < X ≤ x + dx) = f x ( x)dx (7.15)

Bayes’ theorem can be employed in the same fashion for the continuous
case as in the discrete case as follows:
P(ε / θ ) f ' (θ )
f ' ' (θ ) = ∞ (7.16)
∫ P(ε / θ ) f ' (θ )dθ
−∞

where P(ε / θ ) is the conditional probability of obtaining a particular

experimental outcome, ε, assuming that the parameter of interest is equal to θ.


Since P(ε / θ ) is a function of θ it is also called the likelihood function of θ and is

denoted as L(θ). Noting that the denominator in Equation 7.16 is independent of


θ, f ' ' (θ ) can also be expressed as:
f ' ' (θ ) = kL(θ ) f ' (θ ) (7.17)

162
where k is a normalizing constant scaling f ' ' (θ ) to become a proper PDF and

L(θ) is the likelihood of observing the experimental outcome ε assuming a given


value, θ.
In most engineering problems, more than one parameter needs to be
characterized. Even in the simplest case of a linear fit, two parameters are needed,
the intercept and slope. As a result, Equation 7.16 always becomes a bit more
complicated because the denominator has to be integrated for each parameter as
follows:
r r r
r P (ε / φ ) f ' (φ )
f ' ' (φ ) = ∞ ∞
(7.18)
r r r
∫ ∫ P(ε / φ ) f ' (φ )dφ1.....dφ n
......
−∞ −∞
r
where a vector of model parameters, φ , needs to be updated based on a number

of observations.
The integration in Equation 7.18 may become extremely troublesome,
especially when a nonlinear model is being calibrated. The moments of the
updated distribution may also become impossible to compute analytically.
Although the Bayesian approach is a powerful tool, the lack of an analytical
solution makes direct application of the Bayesian approach impractical for most
engineering applications, specifically in those cases dealing with nonlinear
problems.

163
7.3 FIRST-ORDER, SECOND-MOMENT BAYESIAN METHOD

An analytical approach called the First-order, Second-moment Bayesian


Method (FSBM) is proposed by Gilbert (1999) to deal with nonlinear problems.
This method utilizes analytical approximations in updating model parameters
based on experimental observations.
Instead of presenting the mathematical description of the FSBM (which
can be found in Gilbert, 1999), an example application is presented in this section
to assist the reader in obtaining a feel for this method. This example application
introduces the methodology of applying the FSBM to calibrate a simple model
that represents the shear wave velocity profile at a geotechnical site.
The assumed shear wave velocity profile model is presented in Equation
7.19. The model relates the mean estimate of shear wave velocity, µ, to depth, z.
The covariance structure, COV(Vsi,Vsj), is formulated to account for magnitude of
the scatter (or deviations from the mean estimate) and correlation between shear
wave velocities at similar depths anticipating a horizontally layered soil deposit.
µ = φ1 + φ 2 z 0.25 (7.19a)
zi − z j
COV (Vsi ,Vsj ) = φ 3 2 exp(−2 ) (7.19b)
φ4

where:
φ1 is the mean shear wave velocity at the surface,
φ 2 is the empirical constant relating mean shear wave velocity to depth,
φ 3 is the standard deviation of shear wave velocity, and
φ 4 is the scale of fluctuation.

164
Based on past experience, the following prior information is generated.
The expected values, µφ, of model parameters are tabulated in Table 7.2. A rather
large coefficient of variation, δφ, is used to represent a low confidence in the
initial guess (prior information). The model parameters are assumed to be
independent from each other and the resulting covariance structure, Cφ, for the
model parameters is shown in Table 7.3.

Table 7.2 Prior information regarding the model parameters in the FSBM
example

Parameter µφ δφ
φ1 50 0.50
φ2 100 0.50
φ3 20 0.50
φ4 5 0.50

Table 7.3 Prior covariance structure of the model parameters in the FSBM
example

Cφ φ1 φ2 φ3 φ4
φ1 625 0 0 0
φ2 0 2500 0 0
φ3 0 0 100 0
φ4 0 0 0 6.25

165
The covariance structure within the model should not be confused with the
covariance structure of the model parameters. The covariance structure within the
model expresses the discrepancy between the shear wave velocity profile and the
mathematical representation (model), and it also acknowledges the
interrelationship between shear wave velocities at similar depths and therefore the
interrelationship between closely spaced measurements (in terms of depth). Most
soil deposits are horizontally layered systems and soils sampled from similar
depths are likely to be taken from the same layer resulting in the engineering
properties of these samples to be similar. On the other hand, the covariance
structure of the model parameters describes the uncertainty associated with the
model parameters and how they are related to each other.
The shear wave velocity profile model discussed above is calibrated using
the data presented in Table 7.4 collected at the geotechnical site of interest using
an in-situ seismic method.

Table 7.4 Data used to calibrate the model parameters in the FSBM example

Depth Shear Wave Velocity


Measurement
(m) (m/s)
1 1.0 125
2 5.0 150
3 5.5 160
4 10.0 200

166
In this application, a multivariate normal likelihood function is used with
FSBM to calibrate the model which can be formulated as:
r r 1  1 r r −1 r r 
L(ε \ φ ) = 1
exp − (ε − µ ε ) T Cε (ε − µ ε ) (7.20)
 2 
Cε 2

where:
ε 1   µ1   ρ1,1σ ε 2 .. .. ρ1,nσ ε 2 
   :   
r : r    : :: :: : 
ε =   , µ ε =   , Cε =  ,
:  :   : :: :: : 
ε n  µ n   ρ n,1σ ε 2 .. 2
.. ρ n,nσ ε 

Cε = the matrix determinant of Cε, and T = superscript denoting the matrix

transpose.
The FSBM is based on using a second-order Taylor series to approximate
the likelihood function. The Taylor series is expanded about a set of values of
r
model parameters, φ * , at which the natural logarithm of the likelihood function,
r
g (φ ) , is maximized as follows:
r
[ r r
g (φ ) = ln L(ε \ φ )] (7.21)
 r r T  ∂g 
r*
{ }
g (φ ) +  φ − φ * 

 

∂φ i φ * 
r  

r

g (φ ) ≡ (7.22)
 r r  2  r r 
2
1
{
+  φ −φ *  }
T ∂ g

∂φ ∂φ

{
 φ −φ
*
}


  i j φ * 
r

The values of model parameters maximizing the natural logarithm of


Equation 7.20 are estimated by evaluating the following matrices:

167
125 
 
r 150 
ε =  (7.23a)
160 
200
 125 − (φ1* + φ 2 * *10.25 ) 
 * * 
r r  150 − (φ1 + φ 2 * 5 0.25 ) 
ε − µε =  * * 0.25 
(7.23b)
160 − (φ1 + φ 2 * 5.5 )
 200 − (φ * + φ * *10 0.25 ) 
 1 2 
 2 2 4 2 4.5 2 9 
 φ3* φ3* exp(−2 * ) φ3* exp(−2 * ) φ3* exp(−2 * ) 
 φ 4 φ 4 φ 4 
φ *2 exp(−2 4 ) φ3*2 *2 0.5 *2
φ3 exp(−2 * ) φ3 exp(−2 * )
5 
 3 (7.23c)
φ4* φ4 φ4 
Cε =  2
4.5 2 0.5 2 2 4.5 
φ3* exp(−2 * ) φ3* exp(−2 * ) φ3* φ3* exp(−2 * )
 φ4 φ4 φ4 
 *2 9 * 2 5 * 2 4 .5 2 
φ3 exp(−2 * ) φ3 exp(−2 * ) φ3 exp(−2 * ) φ3* 
 φ4 φ4 φ4 

After taking the natural logarithms of both sides of Equation 7.17 and
r
substituting Taylor’s approximation for g (φ ) (Equation 7.22), the updated

(posterior) model parameters can be estimated using:


r
[ ] r r
]
µ φ / ε ≡ Cφ / ε Cφ −1 µ φ − [G ' ']φ * + {G '} (7.24a)

[ ]
r
−1
Cφ / ε ≡ Cφ − G ' ' ]−1
(7.24b)
∂g (φ ) r
where: G ' = = vector of first partial derivatives of g (φ ) ,
∂φ i
r
∂ 2 g (φ ) r
G' ' = = matrix of second partial derivatives of g (φ ) ,
∂φ i ∂φ j

Cφ = prior covariance matrix of model parameters,


Cφ / ε = posterior covariance matrix of model parameters,
r
µ φ = prior mean vector of model parameters,
r
µ φ / ε = posterior mean vector of model parameters, and

168
r
φ * = expansion point of Taylor series approximation.

The calibrated model parameters are tabulated in Table 7.5. The


uncertainty in the estimates of all model parameters is observed to decrease. More
improvement is observed in the case of φ2 and φ3 parameters suggesting that the
data have provided more information regarding these parameters. If a
measurement had been performed at the ground surface, a lot of improvement in
the φ1 parameter would have been observed. The scale of fluctuation, φ4, requires
more data points to be evaluated with certainty and as a result little improvement
is observed in this parameter.

Table 7.5 Comparison of the prior and posterior information regarding the
model parameters in the FSBM example

Prior Posterior
Parameter µφ' δφ' µφ'' δφ''
φ1 50 0.50 38.05 0.34
φ2 100 0.50 85.15 0.12
φ3 20 0.50 10.84 0.16
φ4 5 0.50 3.89 0.37

Table 7.6 presents the covariance structure of the model parameters in


light of the experimental observations. The negative correlation between φ1 and
φ2, and the positive correlation between φ3 and φ4 are results of working with the
same limited data set in evaluating the model parameters. As the intercept of the
mean of the profile increases, the slope decreases. Similarly, as the variability
around the mean increases, the scale of fluctuation increases with the variability.
The results do not necessarily indicate a physical correlation between the model
parameters in such an application.

169
Table 7.6 Posterior covariance structure of the model parameters in the FSBM
example

Cφ φ1 φ2 φ3 φ4
φ1 289.81 -191.19 0.00 0.00
φ2 -191.19 139.06 0.00 0.00
φ3 0.00 0.00 10.12 0.98
φ4 0.00 0.00 0.98 3.38

Figure 7.3 illustrates this point with a simple example using a linear fit.
The best linear fit to this data set is: y = 34.875x + 38.506. Another reasonable
linear fit to this data set is expected to have a smaller slope for a larger intercept
(or a larger slope for a smaller intercept). This relationship between the intercept
and slope is a result of working with a limited set of data with certain
characteristics. Thus, an imaginary correlation between model parameters is
observed upon updating prior information based on limited number of
observations.

150

y = 34.875x + 38.506

100

Dependent
Variable, y

50

y = 0.1348x + 60

0
0 0.2 0.4 0.6 0.8 1
Independent Variable, x

Figure 7.3 Imaginary correlation between model parameters upon updating


prior information based on limited number of observations

170
Table 7.7 tabulates the estimates of shear wave velocity at a depth of 10 m
using prior and posterior model parameters. The uncertainties associated with
these estimates are also presented in this table. Upon calibration of the model with
field observations, a tremendous reduction in the uncertainty regarding model
parameters takes place. As a result, the standard deviation associated with the
estimate of shear wave velocity at a depth of 10 m decreases significantly from 95
m/s to 13 m/s. An experienced engineer familiar with the geotechnical site of
interest might have a better idea about the model parameters prior to in-situ
testing and, under such circumstances, the reduction in uncertainty would not be
as significant.

Table 7.7 Posterior covariance structure of the model parameters in the FSBM
example

Prior Posterior
Expected Value of Vs at 10 m 227.83 189.47
Variance of Vs at 10 m 9030.69 177.09
Standard Deviation of Vs at 10 m 95.03 13.31
Variance from Model Uncertainty 8630.69 59.68
Variance from Random Variability 400.00 117.41
Contribution of Model Uncertainty 0.96 0.34
Contribution of Random Variability 0.04 0.66

171
7.4 FORM OF PROPOSED EQUATIONS

A four-parameter (reference strain, curvature coefficient, scaling


coefficient and small-strain material damping ratio) soil model was presented in
Chapter Six. In Chapter Four, the impact of soil type and loading conditions on
nonlinear soil behavior were discussed.
In this section, the problem of accounting for the impact of soil type and
loading conditions on nonlinear soil behavior is structured analytically. The
parameters of the soil model are related to soil plasticity, void ratio, confining
pressure, overconsolidation ratio, loading frequency and number of loading
cycles. The parameters (denoted as φi in these equations) need to be calibrated
based on the experimental observations (resonant column and torsional shear test
results). The resulting equations can then be utilized to estimate the mean
normalized modulus reduction and material damping curves for a given soil type
and loading conditions.
In an effort to represent the trends regarding the scatter of the data, a
second set of equations for standard deviation are proposed in this section. The
modeled correlation structure of the data is also presented herein.

7.4.1 Equations for Mean

Equations used in representing the impact of soil type and loading


conditions on nonlinear soil behavior are described in this section. Proposed
forms of these equations are based on the trends reported in the literature and
experimental observations in the course of this study.

172
As discussed in Section 5.2, the hyperbolic soil model originally proposed
by Hardin and Drnevich (1972b) is slightly modified and this modified
relationship is utilized to represent a normalized modulus reduction curve as:
G 1
= a
(7.25)
Gmax γ 
1 +  
γ r 

where: G/Gmax = normalized modulus,


γ = shearing strain,
γr = reference strain, and
a = curvature coefficient.
The two parameters in this model (γr and a) can be related to soil type and
loading conditions (σo’ and OCR) as follows:
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ
3 4
(7.26a)
a = φ5 (7.26b)

where: σo’ = mean effective confining pressure (atm),


PI = soil plasticity (%),
OCR = overconsolidation ratio, and
φ1 through φ5 = parameters that relate the normalized modulus reduction
curve to soil type and loading conditions.
These equations can model the shifting of the normalized modulus
reduction curve with increasing soil plasticity, overconsolidation and confining
pressure. When subsets of the data are evaluated, the sensitivity of the curvature
coefficient to soil type and loading conditions can also be studied.

173
In Section 5.3, the material damping curve is calculated assuming Masing
behavior. The calculated material damping curve is then scaled and shifted using:
D = F * DMa sin g + Dmin (7.27)

where:
DMa sin g = c1 DMa sin g ,a =1.0 + c 2 DMa sin g ,a =1.0 2 + c3 DMa sin g ,a =1.0 3 ,
 γ +γ r  
 γ − γ r ln  
DMa sin g ,a =1.0
1
= 4  γr  − 2 ,
Π γ2 
 
 γ +γr 

c1 = -1.1143a 2 + 1.8618a + 0.2523 ,


c 2 = 0.0805a 2 - 0.0710a - 0.0095 ,
c3 = - 0.0005a 2 + 0.0002a + 0.0003 ,
0.1
 G 
F = b *   ,
 Gmax 

b = scaling coefficient,
G/Gmax = normalized modulus,
DMasing = damping estimated based on Masing Behavior, and
Dmin = small-strain material damping ratio.
The two parameters in this equation (Dmin and b) can be related to soil type
and loading conditions as follows:
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )] (7.28a)

b = φ11 + φ12 * ln( N ) (7.28b)

where: σo’ = mean effective confining pressure (atm),


PI = soil plasticity (%),
OCR = overconsolidation ratio,

174
frq = loading frequency,
N = number of loading cycles, and
φ6 through φ12 = parameters that relate material damping curve to soil type
and loading conditions.
Since the material damping curve is directly related to the normalized
modulus reduction curve, any shift in the normalized modulus reduction curve
(due to increasing soil plasticity, overconsolidation and confining pressure) is also
captured in the material damping curve. Small-strain trends in material damping
(discussed in Chapter Four) are accounted for by modeling Dmin separately as a
function of soil plasticity, overconsolidation, confining pressure and loading
frequency. Finally, the impact of number of loading cycles on the material
damping curve is captured in the equation for scaling coefficient.

7.4.2 Equations for Standard Deviation and Covariance Structure

Scatter of the data around the mean estimate is modeled considering the
characteristics of the normalized modulus reduction and material damping curves.
In the case of the normalized modulus reduction curve, less scatter around
the mean is expected at small strains (at which G/Gmax is about 1.0) and at rather
large strains (at which G/Gmax is less than 20%). Uncertainty maximizes around
the reference strain (at which G/Gmax is equal to 0.5). To capture these
characteristics, the standard deviation for a given normalized modulus reduction
curve is modeled as:
0.25 (G / Gmax − 0.5) 2
σ NG = exp(φ13 ) + − (7.29)
exp(φ14 ) exp(φ14 )

where: σNG = standard deviation for normalized modulus reduction curve,


175
G/Gmax = estimated normalized shear modulus, and
φ13 and φ14 = parameters that relate standard deviation to mean estimate of
normalized shear modulus.
Figure 7.4 shows the variation of standard deviation with normalized shear
modulus predicted with Equation 7.29. This equation predicts smaller values of
standard deviation for G/Gmax values close to 1.0 and to 0.0. The scatter relative to
an estimated normalized modulus reduction curve is presented in Figure 7.5. This
scatter pattern looks a lot like the lower bound and upper bound curves proposed
by Seed et al. (1986) for sands shown in Figure 4.22.

0.5

0.4

0.3
σNG
0.2

0.1

0.0
0.0 0.2 0.4 0.6 0.8 1.0
G/Gmax

Figure 7.4 Variation of standard deviation with normalized shear modulus

176
1.2

0.8
G/Gmax
0.4

0.0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ , %

Figure 7.5 Standard deviation modeled for normalized modulus reduction curve

Similarly, in case of the material damping curve, less scatter around the
mean is expected at small strains (at which D is close to Dmin) and uncertainty
increases with increasing shearing strain. As a result, standard deviation for
material damping ratio is modeled as follows:
σ D = exp(φ15 ) + exp(φ16 ) * D (7.30)

where: σD = standard deviation for material damping curve,


D = estimated material damping ratio, and
φ15 and φ16 = parameters that relate standard deviation to the mean
estimate of material damping ratio.
Figure 7.6 shows the variation of standard deviation with material
damping ratio predicted with Equation 7.30. This equation predicts larger values
of standard deviation for higher values of D. The scatter relative to an estimated
material damping curve is presented in Figure 7.7. This scatter pattern also looks
like the lower bound and upper bound curves proposed by Seed et al. (1986) for
sands shown in Figure 4.22.

177
4

σD 2

0
0 5 10 15 20 25
D, %

Figure 7.6 Variation of standard deviation with material damping ratio

25

20

15
D,%
10

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ , %

Figure 7.7 Standard deviation modeled for material damping curve

Random variability and correlation between measurements are also


modeled in this study.
Due to equipment and operator errors, limited resolution of test equipment
and electronic noise that reduces measurement quality and precision, the result of
a test may not exactly be reproduced even though another test is performed under
identical circumstances. Measurements performed on the same specimen at the

178
same confining pressure using the same testing method are assumed to be
correlated. This correlation is modeled through using scale of fluctuation about
the mean estimate. Measurements performed at similar strain amplitudes are
modeled to be highly correlated with each other. The covariance structure is
formulated using:
−1 − ln γ i − ln γ j
ρ i , j = exp( ) * exp( ) (7.31)
exp(φ17 ) exp(φ18 )

where: ρi,j = correlation coefficient,


γi = shearing strain at which measurement i is performed,
φ17 = random variability of data, and
φ18 = scale of fluctuation of data about the mean estimate.

7.5 SUMMARY

In this chapter, the First-order, Second-moment Bayesian Method is


introduced. This method is an approximation of Bayes’ theorem, which is a
systematic way of utilizing experience and observations in statistical analysis.
Equations that relate each parameter of the soil model to soil type and
loading conditions are also presented herein. The equations are formulated so that
the impact of strain amplitude, effective isotropic confining pressure, loading
history and loading frequency on dynamic soil behavior can be properly modeled
following the trends discussed in Chapter Four.
Finally, the equations formulated to represent the scatter and covariance
structure of the data are presented near the end of the chapter.

179
CHAPTER 8
STATISTICAL ANALYSIS OF THE RCTS DATA

8.1 INTRODUCTION

The statistical analysis of the RCTS data is performed using the First-
order, Second-moment Bayesian Method (FSBM) discussed in Chapter Seven. A
computer program that utilizes FSBM originally written by Dr. Robert B. Gilbert
is used in the analysis to calibrate the set of equations presented in Section 7.4.
The program was customized for this application through the C++ header and
source files presented in Appendices A through D.
Since the effect of number of cycles, N, and loading frequency, f, on
normalized modulus reduction curve is negligible in the case of the competent
soils that were investigated in the course of this study, the proposed equations in
Section 7.4 ignore the effect of N and f on G/Gmax. As a result, only resonant
column data (which is typically collected over a wider range of shearing strain
than the torsional shear test results) are utilized in the analysis of modulus
reduction. In case of material damping ratio, the proposed equations are calibrated
using the first and tenth cycles of torsional shear data along with the resonant
column data in an effort to model the effect of N and f on material damping ratio.
Since soil type is expected to be one of the most important parameters that
impact nonlinear soil behavior, the data was analyzed in several subsets according
to soil type and geographic location. Table 8.1 presents the distribution of the
specimens within the database according to soil type and geographic location. The

180
soil types can be categorized into four groups: 1) “Clean” Sands (sands with fines
content less than 12%), 2) Sands with High Fines Content (sands with fines
content greater than 12%), 3) Silts, and 4) Clays. The distribution of the
specimens according to these soil groups is tabulated in Table 8.2.

Table 8.1 Distribution of specimens with soil type and geographic location

Geographic Location
Soil Type Northern California Southern California South Carolina Lotung, Taiwan TOTAL
SW-SM 1 3 - - 4
SW-SC 1 - - - 1
SP-SM 2 4 3 - 9
SP 2 - - - 2
SM 5 13 5 2 25
SC-SM 2 2 - - 4
SC 2 1 4 - 7
ML 3 2 - 6 11
MH 2 - 1 - 3
CL-ML 2 2 - - 4
CL 10 16 2 - 28
CH 5 4 3 - 12
TOTAL 37 47 18 8 110

Table 8.2 Distribution of specimens by soil group and geographic location

Geographic Location
Soil Group Soil Type Northern California Southern California South Carolina Lotung, Taiwan TOTAL
SW-SM
"Clean" SW-SC
6 7 3 - 16
Sands SP-SM
SP
Sands with SM
High Fines SC-SM 9 16 9 2 36
Content SC
ML
Silts MH 7 4 1 6 18
CL-ML
CL
Clays 15 20 5 - 40
CH

181
The test results of all specimens from each soil group within a geographic
location were analyzed separately in order to study the effect of geology on model
parameters. Four specimens from South Carolina (specimens UT-39-G, UT-39-
M, UT-39-O, and UT-39-S) were removed from the database following the
analysis because the resonant column results did not follow the general trends
reported in the literature and observed during the course of this study. The
torsional shear results for these soils did follow the general trends but were not of
sufficient strain range to be included. As a result, a second set of analyses was
performed on two soil groups from which data had been discarded. Tables 8.3 and
8.4 present the distribution of the specimens within the database after the four
specimens have been discarded.

Table 8.3 Distribution of specimens with soil type and geographic location for
the updated database

Geographic Location
Soil Type Northern California Southern California South Carolina Lotung, Taiwan TOTAL
SW-SM 1 3 - - 4
SW-SC 1 - - - 1
SP-SM 2 4 3 - 9
SP 2 - - - 2
SM 5 13 4 2 24
SC-SM 2 2 - - 4
SC 2 1 3 - 6
ML 3 2 - 6 11
MH 2 - 1 - 3
CL-ML 2 2 - - 4
CL 10 16 - - 26
CH 5 4 3 - 12
TOTAL 37 47 14 8 106

182
Table 8.4 Distribution of specimens by soil group and geographic location for
the updated database

Geographic Location
Soil Group Soil Type Northern California Southern California South Carolina Lotung, Taiwan TOTAL
SW-SM
"Clean" SW-SC
6 7 3 - 16
Sands SP-SM
SP
Sands with SM
High Fines SC-SM 9 16 7 2 34
Content SC
ML
Silts MH 7 4 1 6 18
CL-ML
CL
Clays 15 20 3 - 38
CH

The test results of all specimens from each soil group (regardless of its
geographic location) were also analyzed in order to study the effect of soil type on
model parameters. These analyses were performed on the updated database (after
discarding test results of the four specimens from South Carolina).
After concluding that the data were being stretched too thin to calibrate the
model using the subsets, the complete database was utilized in the analysis.
Section 8.3 presents the analysis of all credible data (within the updated database),
which forms the basis of the following chapters regarding the predictions based
on the calibrated model.

183
8.2 ANALYSIS OF SUBSETS OF THE DATA

As discussed in Section 8.1, the data were analyzed in several subsets


according to soil type and geographic location. This section presents the results of
these analyses in both graphical and tabular forms. Table 8.5 presents the prior
mean values and variances of the model parameters that were utilized in the
analysis. Prior mean values of the model parameters are initial guesses based on
general trends reported in the literature and observed during the course of this
study. Variances of the model parameters reflect the confidence in these initial
guesses.

8.2.1 Sorted by Location and Soil Group

The analyses of fourteen subsets of the data are presented in this section.
In most cases, a very limited number of points (for a meaningful analysis from a
statistical standpoint) are utilized in the analysis presented herein as a result of
dividing the data into many subsets. Consequently, the results of these analyses
are presented only for qualitative purposes to study the effect of geology on
model parameters.

8.2.1.1 Samples from Northern California

The results of the analysis of the four soil groups from Northern California
are presented in this section. The updated mean values and variances of the model
parameters for the soil groups are tabulated in Table 8.6.

184
Table 8.5 Prior mean values and variances of the model parameters

Sands with High Fines


Model "Clean" Sands Silts Clays
Content
Parameters*
Mean Variance Mean Variance Mean Variance Mean Variance
φ1 3.50E-02 1.00E-04 3.50E-02 1.00E-04 3.50E-02 1.00E-04 3.50E-02 1.00E-04
φ2 1.00E-03 6.25E-06 1.00E-03 6.25E-06 1.00E-03 6.25E-06 1.00E-03 6.25E-06
φ3 2.50E-01 1.00E-02 2.50E-01 1.00E-02 2.50E-01 1.00E-02 2.50E-01 1.00E-02
φ4 5.00E-01 1.00E-02 5.00E-01 1.00E-02 5.00E-01 1.00E-02 5.00E-01 1.00E-02
φ5 8.50E-01 2.25E-02 9.00E-01 2.25E-02 1.00E+00 2.25E-02 1.05E+00 2.25E-02
φ6 8.00E-01 2.50E-01 8.50E-01 2.50E-01 1.00E+00 2.50E-01 1.10E+00 2.50E-01
φ7 1.00E-02 2.50E-05 1.00E-02 2.50E-05 1.00E-02 2.50E-05 1.00E-02 2.50E-05
φ8 -1.00E-01 2.50E-03 -1.00E-01 2.50E-03 -1.00E-01 2.50E-03 -1.00E-01 2.50E-03
φ9 -1.50E-01 1.00E-02 -1.50E-01 1.00E-02 -1.50E-01 1.00E-02 -1.50E-01 1.00E-02
φ10 2.00E-01 1.00E-02 2.00E-01 1.00E-02 2.00E-01 1.00E-02 2.00E-01 1.00E-02
φ11 7.50E-01 1.00E-02 7.00E-01 1.00E-02 6.50E-01 1.00E-02 6.00E-01 1.00E-02
φ12 -3.00E-02 1.00E-04 -2.00E-02 1.00E-04 -1.00E-02 1.00E-04 0.00E+00 1.00E-04
φ13 -5.00E+00 9.00E+00 -5.00E+00 9.00E+00 -5.00E+00 9.00E+00 -5.00E+00 9.00E+00
φ14 4.50E+00 4.00E+00 4.50E+00 4.00E+00 4.50E+00 4.00E+00 4.50E+00 4.00E+00
φ15 -5.00E+00 9.00E+00 -5.00E+00 9.00E+00 -5.00E+00 9.00E+00 -5.00E+00 9.00E+00
φ16 -1.00E+00 1.00E+00 -1.00E+00 1.00E+00 -1.00E+00 1.00E+00 -1.00E+00 1.00E+00
φ17 4.00E+00 9.00E+00 4.00E+00 9.00E+00 4.00E+00 9.00E+00 4.00E+00 9.00E+00
φ18 2.00E+00 1.00E+00 2.00E+00 1.00E+00 2.00E+00 1.00E+00 2.00E+00 1.00E+00
* Model parameters were defined in Equations 7.25 through 7.31 in Section 7.4.

G 1
= a
G max γ 
1 +  
γ r 
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ 3 4

a = φ5
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin
 G max 
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )]

b = φ11 + φ12 * ln( N )

185
Table 8.6 Updated mean values and variances of the model parameters for the
soils from Northern California

Sands with High Fines


Model "Clean" Sands Silts Clays
Content
Parameters*
Mean Variance Mean Variance Mean Variance Mean Variance
φ1 4.26E-02 9.50E-06 2.76E-02 2.05E-06 3.97E-02 8.49E-06 3.39E-02 3.65E-05
φ2 -2.28E-03 1.20E-07 -3.42E-05 8.90E-09 -3.27E-06 4.25E-08 1.75E-03 5.30E-08
φ3 2.50E-01 1.00E-02 2.50E-01 1.00E-02 2.50E-01 1.00E-02 2.97E-01 4.62E-03
φ4 2.53E-01 1.90E-03 2.49E-01 1.26E-03 1.50E-01 1.41E-03 2.78E-01 1.07E-03
φ5 9.41E-01 6.28E-04 8.83E-01 2.33E-04 1.05E+00 8.37E-04 9.93E-01 4.66E-04
φ6 9.95E-01 1.25E-02 1.26E+00 2.19E-02 9.54E-01 6.99E-03 1.05E+00 1.30E-02
φ7 1.24E-02 2.37E-05 1.61E-02 2.21E-05 8.04E-03 1.95E-05 8.20E-04 1.05E-05
φ8 -1.00E-01 2.50E-03 -1.00E-01 2.50E-03 -9.95E-02 2.50E-03 -1.01E-01 2.50E-03
φ9 1.25E-01 5.15E-03 -1.73E-01 3.56E-03 -2.30E-01 2.40E-03 -1.72E-01 1.67E-03
φ10 2.85E-01 3.37E-03 2.19E-01 2.09E-03 2.90E-01 3.59E-03 3.23E-01 2.98E-03
φ11 7.50E-01 1.96E-03 6.71E-01 1.09E-03 4.11E-01 6.93E-04 4.40E-01 5.99E-04
φ12 -2.83E-02 3.40E-05 -1.50E-02 3.28E-05 1.58E-02 1.81E-05 2.33E-02 1.61E-05
φ13 -3.70E+00 7.49E-02 -3.35E+00 1.76E-02 -5.00E+00 9.00E+00 -5.28E+00 1.27E-01
φ14 4.71E+00 3.07E-01 5.11E+00 3.50E+00 4.31E+00 6.40E-02 3.89E+00 4.45E-02
φ15 -5.03E+00 8.97E+00 -5.06E+00 8.90E+00 -5.01E+00 8.99E+00 -5.22E+00 8.75E+00
φ16 -5.02E-01 1.39E-02 -4.29E-01 1.14E-02 -9.62E-01 1.01E-02 -6.70E-01 6.16E-03
φ17 4.85E+00 2.14E-01 3.79E+00 9.24E-02 4.03E+00 8.98E+00 7.04E+00 6.66E-01
φ18 2.95E+00 6.66E-02 2.67E+00 5.89E-02 1.94E+00 4.60E-02 2.18E+00 2.52E-02
* Model parameters were defined in Equations 7.25 through 7.31 in Section 7.4.

G 1
= a
G max γ 
1 +  
γ r 
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ 3 4

a = φ5
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin
 G max 
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )]

b = φ11 + φ12 * ln( N )

186
Since most of the specimens from Northern California are normally
consolidated, the updated mean values and variances of φ3 and φ8 (which
represent the effect of overconsolidation ratio on reference strain and small-strain
material damping ratio, respectively) are almost identical to prior values. Hence,
the data have not provided much information regarding these parameters. The
values of φ1 (which represents the reference strain of a nonplastic soil at 1 atm
confining pressure), φ4 (which represents the effect of confining pressure on
reference strain), φ6 (which represents the small-strain material damping ratio of a
nonplastic soil at 1 atm confining pressure deformed at 1 Hz loading frequency)
and φ10 (which represents the effect of loading frequency on small-strain material
damping ratio) are observed to be consistent between soil groups.
The comparisons of the measurements with the predicted values based on
the calibrated models for the four soil groups are presented in Figures 8.1 through
8.4. Significantly less error is observed in the prediction of normalized shear
modulus than in the prediction of material damping predictions ratio for all soil
groups. This difference can be attributed to material damping ratio being sensitive
to the characteristics of the complete stress-strain loop while normalized shear
modulus is only related to the end points of the stress-strain loop. Consequently,
measurement and prediction of material damping ratio is more complicated than
measurement and prediction of normalized shear modulus.

187
1.2 25
Predicted Normalized Modulus

Predicted Material Damping


(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.1 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for “clean” sands from
Northern California

1.2 25
Predicted Normalized Modulus

Predicted Material Damping

(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.2 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for sands with high fines
content from Northern California

188
1.2 25
Predicted Normalized Modulus

Predicted Material Damping


(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.3 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for silts from Northern
California

1.2 25
Predicted Normalized Modulus

Predicted Material Damping

(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.4 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for clays from Northern
California

189
8.2.1.2 Samples from Southern California

The updated mean values and variances of the model parameters for the
four soil groups from Southern California are tabulated in Table 8.7. As in the
case of Northern California, the data have not provided much information
regarding the φ3 and φ8 parameters. The φ1, φ4, φ6 and φ10 parameters are again
observed to be consistent between soil groups. φ5 (which is the curvature
coefficient) is observed to slightly increase with decreasing particle size. The
comparisons of the measured and predicted values based on the calibrated models
are presented in Figures 8.5 through 8.8.

8.2.1.3 Samples from South Carolina

8.2.1.3.1 Analysis of Test Results from All (Eighteen) Specimens

The updated mean values and variances of the model parameters for the
four soil groups from South Carolina are tabulated in Table 8.8. The model
parameters are observed to be extremely inconsistent. Part of the problem is
believed to be the result of analyzing a very small dataset in the case of “clean”
sands and silts.
The comparisons of the measured and predicted values based on the
calibrated models are presented in Figures 8.9 through 8.12. Test results from four
specimens do not agree with the observed trends and reduce the quality of the
predictions for two soil groups, sands with high fines content and clays. As a
result, these specimens were identified and discarded from the database.

190
Table 8.7 Updated mean values and variances of the model parameters for the
soils from Southern California

Sands with High Fines


Model "Clean" Sands Silts Clays
Content
Parameters*
Mean Variance Mean Variance Mean Variance Mean Variance
φ1 2.53E-02 6.44E-06 3.51E-02 5.30E-06 5.18E-02 4.53E-05 3.52E-02 5.76E-06
φ2 1.00E-03 6.25E-06 1.34E-03 8.41E-08 5.96E-05 9.19E-07 7.07E-04 9.12E-09
φ3 2.50E-01 1.00E-02 2.62E-01 9.44E-03 2.50E-01 1.00E-02 3.69E-01 7.30E-03
φ4 4.62E-01 2.97E-03 5.04E-01 1.35E-03 4.26E-01 9.34E-04 2.97E-01 5.31E-04
φ5 8.34E-01 6.37E-04 8.58E-01 3.62E-04 9.40E-01 1.41E-03 9.50E-01 2.25E-04
φ6 8.42E-01 3.56E-02 8.26E-01 1.19E-02 7.75E-01 7.47E-03 1.01E+00 4.39E-03
φ7 1.00E-02 2.50E-05 1.29E-02 2.36E-05 8.55E-03 2.41E-05 3.92E-04 1.06E-05
φ8 -1.00E-01 2.50E-03 -1.00E-01 2.50E-03 -1.00E-01 2.50E-03 -1.01E-01 2.49E-03
φ9 -2.90E-01 6.32E-03 -4.18E-01 2.99E-03 -1.30E-01 1.50E-03 -1.97E-01 9.35E-04
φ10 2.53E-01 5.59E-03 2.37E-01 2.70E-03 1.97E-01 2.90E-03 3.74E-01 1.76E-03
φ11 7.62E-01 2.08E-03 7.70E-01 1.61E-03 6.67E-01 2.24E-03 5.18E-01 4.84E-04
φ12 -2.67E-02 4.19E-05 -2.41E-02 3.12E-05 -1.95E-02 3.56E-05 1.68E-02 1.18E-05
φ13 -5.05E+00 8.93E+00 -5.00E+00 9.00E+00 -5.02E+00 8.98E+00 -5.68E+00 3.93E-02
φ14 4.51E+00 6.15E-02 3.27E+00 3.49E-02 4.92E+00 8.97E-02 4.39E+00 1.98E-02
φ15 -5.00E+00 9.00E+00 -5.04E+00 8.96E+00 -5.00E+00 9.00E+00 -5.03E+00 8.97E+00
φ16 -5.97E-01 1.38E-02 -1.81E-01 7.62E-03 -1.23E+00 1.54E-02 -8.40E-01 2.98E-03
φ17 3.99E+00 9.00E+00 4.00E+00 9.00E+00 4.06E+00 8.93E+00 4.23E+00 8.80E+00
φ18 2.19E+00 5.19E-02 2.80E+00 2.65E-02 1.53E+00 8.08E-02 2.08E+00 1.31E-02
* Model parameters were defined in Equations 7.25 through 7.31 in Section 7.4.

G 1
= a
G max γ 
1 +  
γ r 
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ 3 4

a = φ5
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin
 G max 
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )]

b = φ11 + φ12 * ln( N )

191
1.2 25
Predicted Normalized Modulus

Predicted Material Damping


(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.5 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for “clean” sands from
Southern California

1.2 25
Predicted Normalized Modulus

Predicted Material Damping

(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.6 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for sands with high fines
content from Southern California

192
1.2 25
Predicted Normalized Modulus

Predicted Material Damping


(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.7 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for silts from Southern
California

1.2 25
Predicted Normalized Modulus

Predicted Material Damping

(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.8 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for clays from Southern
California

193
Table 8.8 Updated mean values and variances of the model parameters for the
soils from South Carolina

Sands with High Fines


Model "Clean" Sands Silts Clays
Content
Parameters*
Mean Variance Mean Variance Mean Variance Mean Variance
φ1 6.83E-02 3.33E-05 3.22E-02 2.32E-05 3.47E-02 9.99E-05 3.20E-02 9.90E-05
φ2 5.82E-03 5.21E-06 -5.03E-04 7.03E-08 9.34E-04 9.82E-09 2.79E-03 2.90E-07
φ3 2.50E-01 1.00E-02 2.50E-01 1.00E-02 2.83E-01 9.88E-03 2.50E-01 1.00E-02
φ4 -3.88E-01 2.85E-03 -3.80E-02 3.52E-03 2.30E-01 1.98E-03 7.04E-02 1.92E-03
φ5 8.38E-01 6.73E-03 8.96E-01 1.87E-03 1.17E+00 2.18E-03 1.09E+00 5.85E-03
φ6 1.57E-01 5.13E-05 2.82E-01 1.47E-03 5.13E-01 1.40E-01 -2.18E-01 8.59E-02
φ7 1.01E-02 2.49E-05 1.68E-02 1.00E-05 4.90E-03 9.65E-06 5.70E-03 1.24E-05
φ8 -1.00E-01 2.50E-03 -9.87E-02 2.50E-03 -1.12E-01 2.10E-03 -1.00E-01 2.50E-03
φ9 -1.46E-01 6.68E-03 4.71E-02 6.16E-03 -1.25E-01 4.22E-03 -2.93E-01 7.83E-03
φ10 2.27E-01 1.26E-03 3.05E-01 8.95E-03 9.62E-02 8.47E-04 1.42E-01 5.69E-03
φ11 8.87E-01 2.36E-03 6.68E-01 3.05E-03 4.35E-01 1.12E-03 3.53E-01 4.25E-03
φ12 -1.26E-02 3.76E-05 -2.05E-02 4.69E-05 2.32E-02 3.22E-05 -2.08E-02 7.07E-05
φ13 -6.57E+00 6.41E-01 -5.04E+00 8.97E+00 -5.84E+00 3.42E-01 -4.65E+00 1.01E-01
φ14 5.38E+00 2.88E-01 2.44E+00 7.23E-02 5.98E+00 3.14E-01 3.94E+00 2.88E-01
φ15 -5.00E+00 9.00E+00 -5.00E+00 9.00E+00 -5.02E+00 8.98E+00 -5.00E+00 9.00E+00
φ16 -1.80E+00 1.43E-02 -1.20E-01 1.75E-02 -1.59E+00 1.11E-02 8.52E-01 3.96E-02
φ17 3.36E+00 5.19E-01 4.00E+00 9.00E+00 4.01E+00 8.99E+00 4.03E+00 8.98E+00
φ18 5.66E-01 8.76E-02 2.39E+00 5.09E-02 -3.42E-03 8.17E-02 2.82E+00 1.00E-01
* Model parameters were defined in Equations 7.25 through 7.31 in Section 7.4.

G 1
= a
G max γ 
1 +  
γ r 
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ 3 4

a = φ5
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin
 G max 
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )]

b = φ11 + φ12 * ln( N )

194
1.2 25
Predicted Normalized Modulus

Predicted Material Damping


(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.9 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for “clean” sands from
South Carolina

1.2
Predicted Normalized Modulus

25
Predicted Material Damping

(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.10 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for sands with high fines
content from South Carolina

195
1.2 25
Predicted Normalized Modulus

Predicted Material Damping


(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.11 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for silts from South Carolina

1.2 25
Predicted Normalized Modulus

Predicted Material Damping

(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.12 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for clays from South
Carolina

196
8.2.1.3.2 Analysis of Test Results from Fourteen Specimens

After discarding the four specimens that were identified not to be


consistent with the observed trends, the two soil groups (sands with high fines
content and clays) that were affected by the process of changing the contents of
the database were reevaluated. Considerable improvement in predicted values is
achieved. The updated mean values and variances of the model parameters for
these two soil groups from South Carolina are tabulated in Table 8.9 and the
comparisons of the measured and predicted values based on the calibrated models
are presented in Figures 8.13 through 8.14.

8.2.1.4 Samples from Lotung, Taiwan

The updated mean values and variances of the model parameters for the
two soil groups from Lotung, Taiwan are tabulated in Table 8.10. Inconsistencies
between the model parameters are attributed to analyzing a very small dataset in
the case of sands with high fines content. The comparisons of the measured and
predicted values based on the calibrated models are presented in Figures 8.15 and
8.16.

197
Table 8.9 Updated mean values and variances of the model parameters for the
South Carolina soil groups affected by change in the contents of the
database

Sands with High Fines


Model Clays
Content
Parameters*
Mean Variance Mean Variance
φ1 2.88E-02 7.22E-06 3.25E-02 9.87E-05
φ2 -3.62E-05 7.87E-09 2.64E-03 6.85E-08
φ3 2.48E-01 1.00E-02 2.50E-01 1.00E-02
φ4 6.36E-01 4.38E-03 1.89E-02 1.36E-03
φ5 8.78E-01 7.74E-04 1.29E+00 3.12E-03
φ6 4.37E-01 1.85E-03 9.66E-01 3.93E-02
φ7 1.23E-02 1.56E-06 1.08E-02 1.06E-05
φ8 -8.35E-02 2.43E-03 -1.00E-01 2.50E-03
φ9 -7.45E-02 7.99E-03 -3.84E-02 2.43E-03
φ10 3.25E-01 5.92E-03 2.89E-01 1.59E-03
φ11 7.92E-01 2.54E-03 5.09E-01 2.98E-03
φ12 -3.66E-02 3.96E-05 6.75E-03 5.86E-05
φ13 -5.02E+00 8.98E+00 -5.07E+00 1.59E-01
φ14 4.31E+00 5.55E-02 4.93E+00 2.33E-01
φ15 -5.00E+00 9.00E+00 -5.00E+00 9.00E+00
φ16 -6.92E-01 1.11E-02 -1.62E+00 1.79E-02
φ17 4.00E+00 8.98E+00 3.55E+00 1.09E+00
φ18 1.93E+00 4.71E-02 1.40E+00 1.91E-01
* Model parameters were defined in Equations 7.25
through 7.31 in Section 7.4.

G 1
= a
G max γ 
1 +  
γ r 
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ 3 4

a = φ5
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin
 G max 
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )]

b = φ11 + φ12 * ln( N )

198
1.2
Predicted Normalized Modulus
25

Predicted Material Damping


(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.13 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for sands with high fines
content from South Carolina (After Discarding Specimens UT-39-G
and UT-39-M)

1.2 25
Predicted Normalized Modulus

Predicted Material Damping

(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.14 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for clays from South
Carolina (After Discarding Specimens UT-39-O and UT-39-S)

199
Table 8.10 Updated mean values and variances of the model parameters for the
soils from Lotung, Taiwan

Sands with High Fines


Model Silts
Content
Parameters*
Mean Variance Mean Variance
φ1 2.90E-02 9.78E-06 4.52E-02 1.08E-05
φ2 1.00E-03 6.25E-06 1.69E-04 1.78E-07
φ3 2.50E-01 1.00E-02 2.50E-01 1.00E-02
φ4 5.02E-01 8.88E-03 1.53E-01 1.97E-03
φ5 8.02E-01 6.48E-04 1.02E+00 1.08E-03
φ6 5.46E-01 1.48E-02 5.73E-01 4.88E-03
φ7 1.00E-02 2.50E-05 1.05E-02 2.17E-05
φ8 -1.00E-01 2.50E-03 -1.00E-01 2.50E-03
φ9 -1.64E-01 9.79E-03 -1.39E-01 5.10E-03
φ10 2.37E-01 7.16E-03 1.02E-01 2.83E-03
φ11 8.07E-01 3.16E-03 7.22E-01 2.07E-03
φ12 -3.84E-02 5.39E-05 -1.70E-02 2.89E-05
φ13 -7.12E+00 4.55E+00 -5.01E+00 8.99E+00
φ14 4.79E+00 3.14E-01 4.56E+00 5.02E-02
φ15 -5.00E+00 9.00E+00 -5.01E+00 8.99E+00
φ16 -8.14E-01 5.27E-02 -9.93E-01 8.88E-03
φ17 3.99E+00 8.97E+00 4.00E+00 9.00E+00
φ18 2.74E+00 2.01E-01 9.02E-01 4.99E-02
* Model parameters were defined in Equations 7.25
through 7.31 in Section 7.4.

G 1
= a
G max γ 
1 +  
γ r 
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ 3 4

a = φ5
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin
 G max 
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )]

b = φ11 + φ12 * ln( N )

200
1.2 25
Predicted Normalized Modulus

Predicted Material Damping


(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.15 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for sands with high fines
content from Lotung, Taiwan

1.2 25
Predicted Normalized Modulus

Predicted Material Damping

(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.16 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for silts from Lotung,
Taiwan

201
8.2.1.5 Comparison of the Nonlinear Behavior of Soils from Different
Locations

In order to evaluate the impact of geology on nonlinear soil behavior, the


normalized modulus reduction and material damping curves for a given soil type
under given loading conditions are predicted utilizing the updated model
parameters for different geographic locations. Figures 8.17 through 8.19 present
the comparison of these predictions.
In Figure 8.17, predicted normalized modulus reduction and material
damping curves for a silty sand are shown. The soil is selected to be nonplastic
and normally consolidated in order to analyze a representative material within the
database. The confining pressure is selected to be 1 atm in the same fashion. Ten
cycles of loading at 1 Hz is chosen so that the loading conditions represent the
characteristics of an earthquake. In this figure, the effect of geographic location
(and geology for that matter) on dynamic soil behavior is observed to be
negligible.
Figure 8.18 shows the comparison of predicted nonlinear soil behavior for
a moderate plasticity silt. The characteristics of the soil are again selected so that
a representative material within the database is evaluated. Loading conditions are
selected to be the same as those in the case of the silty sand. In Figure 8.18a,
normalized modulus reduction curves for the two predictions are observed to be
quite similar. On the other hand, the predicted material damping ratios shown in
Figure 8.18b are somewhat different at shearing strains above 0.1 %. This
divergence can be investigated in future studies using a larger database.

202
1.2
(a)
1.0

0.8
0.6
G/Gmax
0.4 Silty Sand from Northern California
0.2 Silty Sand from Southern California
Silty Sand from South Carolina
0.0

25
PI = 0 %
20 OCR = 1
f = 1 Hz
15 N =10
D, % σo' = 1 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 8.17 (a) Normalized modulus reduction and (b) material damping curves
estimated for a nonplastic silty sand using updated mean values of
model parameters calibrated at different geographic locations

203
1.2
(a)
1.0

0.8
0.6
G/Gmax
0.4
Silt from Northern California
0.2 Silt from Lotung, Taiwan
0.0

25
PI = 15 %
20 OCR = 1
f = 1 Hz
15 N =10
D, % σo' = 1 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 8.18 (a) Normalized modulus reduction and (b) material damping curves
estimated for a moderate plasticity silt using updated mean values of
model parameters calibrated at different geographic locations

204
Figure 8.19 shows the comparison of predicted nonlinear soil behavior for
a moderate plasticity clay evaluated for identical loading conditions. As in the
case of the silty sand, the effect of geographic location and geology on dynamic
soil behavior is observed to be negligible.

1.2
(a)
1.0

0.8
0.6
G/Gmax
0.4
Clay from Northern California
0.2 Clay from Southern California
0.0

25
PI = 15 %
20 OCR = 1
f = 1 Hz
15 N =10
D, % σo' = 1 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 8.19 (a) Normalized modulus reduction and (b) material damping curves
estimated for a moderate plasticity clay using updated mean values
of model parameters calibrated at different geographic locations

205
Comparison of the results for the subsets of the data sorted according to
geographic location does not indicate a strong correlation between geology and
nonlinear soil behavior. As a result, soils within a soil group from different
geographic locations shall be analyzed together in the following sections.

8.2.2 Sorted by Soil Group

Since the effect of geology on nonlinear soil behavior is observed not to


be very significant, soils from different geographic locations are combined into
larger datasets. This section presents the results of the analysis of all soils within
each soil group. Since four specimens from South Carolina were previously
identified to be inconsistent with the general trends, the data associated with these
specimens are not included in these analyses.
Table 8.11 presents the updated mean values and variances of the model
parameters calibrated for the four soil groups. Most of the model parameters are
observed to be consistent between soil groups. The value of φ5 (which is the
curvature coefficient) is observed to be slightly different for the coarse grained
and fine grained soils. This trend is consistent with the observations documented
by Hardin and Drnevich (1972b). The values of φ2 and φ7 (which represent the
effect of plasticity on reference strain and small-strain material damping ratio,
respectively) are observed to be quite different for the coarse grained and fine
grained soils. This difference is believed to result from the smaller range in
plasticity and the fewer number of plastic soils sampled in the case of coarse
grained soils. The comparisons of the measured and predicted values based on the
calibrated models are presented in Figures 8.20 through 8.23.

206
Table 8.11 Updated mean values and variances of the model parameters for the
four soil groups

Sands with High Fines


Model "Clean" Sands Silts Clays
Content
Parameters*
Mean Variance Mean Variance Mean Variance Mean Variance
φ1 4.74E-02 9.62E-06 3.34E-02 2.06E-06 4.16E-02 5.18E-06 2.58E-02 5.68E-06
φ2 -2.34E-03 1.63E-07 -5.79E-05 8.09E-09 6.89E-04 7.74E-09 1.95E-03 1.84E-08
φ3 2.50E-01 1.00E-02 2.49E-01 9.94E-03 3.21E-01 7.56E-03 9.92E-02 1.64E-03
φ4 2.34E-01 1.08E-03 4.82E-01 7.46E-04 2.80E-01 8.63E-04 2.26E-01 3.48E-04
φ5 8.95E-01 4.30E-04 8.45E-01 1.49E-04 1.00E+00 4.10E-04 9.75E-01 1.60E-04
φ6 6.88E-01 7.82E-03 8.89E-01 5.86E-03 7.12E-01 3.55E-03 9.58E-01 2.93E-03
φ7 1.22E-02 2.43E-05 2.02E-02 1.91E-05 3.03E-03 2.65E-06 5.65E-03 2.79E-06
φ8 -1.00E-01 2.50E-03 -1.00E-01 2.50E-03 -1.00E-01 2.50E-03 -1.00E-01 2.50E-03
φ9 -1.27E-01 4.00E-03 -3.72E-01 1.83E-03 -1.89E-01 1.95E-03 -1.96E-01 5.21E-04
φ10 2.88E-01 3.14E-03 2.33E-01 1.35E-03 2.34E-01 2.60E-03 3.68E-01 1.19E-03
φ11 7.67E-01 1.59E-03 7.76E-01 7.71E-04 5.92E-01 8.09E-04 4.66E-01 2.69E-04
φ12 -2.83E-02 2.79E-05 -2.94E-02 1.70E-05 -7.67E-04 1.61E-05 2.23E-02 7.13E-06
φ13 -4.14E+00 4.17E-02 -3.98E+00 1.82E-02 -5.02E+00 8.98E+00 -5.65E+00 3.37E-02
φ14 3.61E+00 5.97E-02 4.32E+00 3.30E-02 3.93E+00 2.47E-02 4.00E+00 1.21E-02
φ15 -5.15E+00 8.80E+00 -5.34E+00 8.55E+00 -5.20E+00 8.76E+00 -5.00E+00 9.00E+00
φ16 -2.32E-01 7.56E-03 -2.66E-01 3.40E-03 -6.42E-01 4.78E-03 -7.25E-01 1.92E-03
φ17 5.15E+00 6.91E-02 4.92E+00 3.74E-02 4.06E+00 8.96E+00 7.67E+00 3.51E-01
φ18 3.12E+00 2.88E-02 2.68E+00 1.38E-02 1.94E+00 1.98E-02 2.16E+00 8.08E-03
* Model parameters were defined in Equations 7.25 through 7.31 in Section 7.4.

G 1
= a
G max γ 
1 +  
γ r 
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ 3 4

a = φ5
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin
 G max 
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )]

b = φ11 + φ12 * ln( N )

207
1.2
Predicted Normalized Modulus
25

Predicted Material Damping


(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.20 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for “clean” sands

1.2 25
Predicted Normalized Modulus

Predicted Material Damping

(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.21 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for sands with high fines
content

208
1.2
Predicted Normalized Modulus
25

Predicted Material Damping


(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.22 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for silts

1.2 25
Predicted Normalized Modulus

Predicted Material Damping

(a) (b)
1.0 20
0.8
15
0.6
10
0.4

0.2 5

0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 5 10 15 20 25

Measured Normalized Modulus Measured Material Damping

Figure 8.23 Comparisons of the measured and predicted values of (a) normalized
modulus and (b) material damping ratio for clays

209
In order to evaluate the impact of soil type on nonlinear soil behavior, the
normalized modulus reduction and material damping curves for a given soil under
given loading conditions are predicted utilizing the updated model parameters for
different soil groups. Figure 8.24 presents the comparison of these predictions.
The coarse grained soils are selected to be nonplastic and the fine grained
soils are selected to be of moderate plasticity in order to analyze a representative
material within each soil group. The confining pressure is selected to be 1 atm for
the same reason. Ten cycles of loading at 1 Hz is again chosen so that the loading
conditions represent the characteristics of an earthquake.
In Figure 8.24, the difference in the nonlinear behavior of the soils from
the four soil groups is observed to be small. From a qualitative standpoint, “clean”
sands are observed to be relatively linear (normalized modulus reduction and
material damping curves located at higher strain amplitudes) compared to sands
with high fines content. This trend is consistent with the discrepancy between
normalized modulus reduction and material damping curves reported for
uniformly graded sand specimens (Iwasaki et al., 1978; Kokusho, 1980; and Ni,
1987) and for natural materials (Seed et al., 1986; Sun et al., 1988; Vucetic and
Dobry, 1991; Hwang, 1997; and Darendeli et al., 2001).
The comparison of the predictions in Figure 8.24 indicates that fines
content (the soil group) does not have a very significant impact on nonlinear soil
behavior. Thus, a model calibrated for the complete data set can be successfully
utilized in developing a new family of normalized modulus reduction and material
damping curves.

210
1.2
(a)
1.0

0.8
0.6 "Clean" Sand (PI = 0 %)
G/Gmax Sand with High
0.4 Fines Content (PI = 0 %)
0.2 Silt (PI = 15 %)
Clay (PI = 15 %)
0.0

25

20 OCR = 1
f = 1 Hz
15 N =10
D, % σo' = 1 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 8.24 (a) Normalized modulus reduction and (b) material damping curves
estimated using updated mean values of model parameters calibrated
for different soil groups

211
8.3 ANALYSIS OF ALL CREDIBLE DATA

The predictions based on the calibrated models from the subsets of the
database indicate that the effects of geology (analyzed through geographic
location) and fines content (analyzed through soil groups) on the model
parameters are not very pronounced. As a result, all credible data (after removal
of the four specimens from South Carolina from the database) are analyzed as one
complete data set herein. The recommended values of the model parameters and
recommended nonlinear curves discussed in the following chapters are based on
the analysis presented in this section. As discussed in Section 8.1, only resonant
column data are utilized in the analysis of normalized shear modulus, and, in an
effort to model the effect of N and f on material damping, first and tenth cycles of
torsional shear data along with the resonant column data are utilized in the
analysis of material damping ratio. All credible data used in the analysis are
presented in Figure 8.25.
Table 8.12 presents the prior and the updated mean values and variances
of the model parameters calibrated for all the credible data presented in Figure
8.25. The table indicates considerable reduction in uncertainty (in the form of
variance) in the model parameters. The only exceptions are parameters φ8 (which
represents the impact of overconsolidation ratio on small-strain material damping
ratio) and φ15 (which represents the scatter of material damping ratio at small-
strains). In the case of these two parameters, very little information is gathered
indicating that the quality of the predictions (illustrated in Figures 8.26 and 8.27)
associated with the calibrated model is not very sensitive to these parameters.

212
1.2
(a)
1.0

0.8
0.6
G/Gmax
0.4
0.2 Measured Normalized Shear Modulus
0.0

25
Measured Material Damping Ratio (b)
20

15
D, %
10

0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 8.25 All credible (a) normalized modulus data from the resonant column
tests, and (b) material damping data from the resonant column and
torsional shear tests utilized to calibrate the model parameters.

213
Table 8.12 Comparison of the prior and updated mean values and variances of
the model parameters for all the credible data

Model Prior Updated


Parameters* Mean Variance Mean Variance
φ1 3.50E-02 1.00E-04 3.52E-02 9.99E-07
φ2 1.00E-03 6.25E-06 1.01E-03 4.16E-09
φ3 2.50E-01 1.00E-02 3.25E-01 2.85E-03
φ4 5.00E-01 1.00E-02 3.48E-01 2.20E-04
φ5 9.00E-01 2.25E-02 9.19E-01 6.78E-05
φ6 8.50E-01 2.50E-01 8.01E-01 1.73E-03
φ7 1.00E-02 2.50E-05 1.29E-02 3.82E-06
φ8 -1.00E-01 2.50E-03 -1.07E-01 2.49E-03
φ9 -1.50E-01 1.00E-02 -2.89E-01 4.96E-04
φ10 2.00E-01 1.00E-02 2.92E-01 7.66E-04
φ11 7.00E-01 1.00E-02 6.33E-01 2.23E-04
φ12 -2.00E-02 1.00E-04 -5.66E-03 5.02E-06
φ13 -5.00E+00 9.00E+00 -4.23E+00 5.38E-03
φ14 4.50E+00 4.00E+00 3.62E+00 7.05E-03
φ15 -5.00E+00 9.00E+00 -5.00E+00 9.00E+00
φ16 -1.00E+00 1.00E+00 -2.50E-01 1.06E-03
φ17 4.00E+00 9.00E+00 5.62E+00 1.53E-02
φ18 2.00E+00 1.00E+00 2.78E+00 3.86E-03
* Model parameters were defined in Equations 7.25 through
7.31 in Section 7.4.

G 1
= a
G max γ 
1 +  
γ r 
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ 3 4

a = φ5
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin
 G max 
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )]

b = φ11 + φ12 * ln( N )

214
1.2

1.0
Predicted Normalized Modulus

0.8

0.6

0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2

Measured Normalized Modulus

Figure 8.26 Comparisons of the measured and predicted values of normalized


modulus for all credible data

215
25

20
Predicted Material Damping

15

10

0
0 5 10 15 20 25

Measured Material Damping

Figure 8.27 Comparisons of the measured and predicted values of material


damping for all credible data

216
Table 8.13 shows the covariance structure of the model parameters. This

information can be utilized in calculating model uncertainty associated with the

normalized modulus reduction and material damping curves estimated based on

this model as discussed in Chapter Eleven. Since the model is calibrated to a

rather large database, most of the uncertainty however results from the variability

within the database modeled using parameters φ13 through φ18 (Section 7.4.2).

8.4 SUMMARY

In order to study the effect of soil type and geology, the data was first

analyzed in several subsets according to soil group (“clean” sands, sands with
high fines content, silts and clays) and geographic location (Northern California,

Southern California, South Carolina and Lotung, Taiwan).

Four specimens from South Carolina (specimens UT-39-G, UT-39-M,

UT-39-O, and UT-39-S) were removed from the database following the analysis

because the resonant column results did not follow the general trends reported in
the literature and observed during the course of this study. The torsional shear

results for these specimens did follow the general trends but were not of sufficient

strain range to be included. At this point, the analyses affected by the change in
the content of the database were repeated and the rest of the analyses were carried

out without utilizing the data associated with these four specimens.

The test results of all specimens from each soil group (regardless of

geographic location) were also analyzed in order to study the effect of soil type on

model parameters.

217
Table 8.13 Covariance structure of the updated model parameters for all the
credible data

φi* φ1 φ2 φ3 φ4 φ5 φ6 φ7 φ8 φ9 φ10 φ11 φ12 φ13 φ14 φ15 φ16 φ17 φ18
φ1 1.00 -0.10 0.03 -0.39 -0.25 -0.05 0.05 0.00 0.05 0.03 0.51 -0.10 -0.06 0.03 0.00 -0.07 -0.06 -0.12
φ2 -0.10 1.00 -0.55 -0.27 -0.10 0.03 -0.18 0.00 0.06 0.05 0.15 0.06 0.03 0.00 0.00 -0.04 0.01 -0.04
φ3 0.03 -0.55 1.00 0.24 -0.01 -0.06 0.12 0.00 -0.02 -0.01 0.00 -0.01 -0.03 -0.03 0.00 0.03 -0.01 0.02
φ4 -0.39 -0.27 0.24 1.00 -0.01 0.01 0.05 0.00 -0.23 -0.03 -0.09 0.08 0.00 -0.13 0.00 0.14 0.03 0.14
φ5 -0.25 -0.10 -0.01 -0.01 1.00 -0.06 -0.02 0.00 -0.05 0.02 -0.51 0.00 0.11 -0.17 0.00 0.12 0.02 0.09
φ6 -0.05 0.03 -0.06 0.01 -0.06 1.00 -0.22 -0.01 -0.20 -0.56 0.24 -0.12 -0.02 0.08 0.00 -0.30 0.04 -0.10
φ7 0.05 -0.18 0.12 0.05 -0.02 -0.22 1.00 0.00 -0.18 -0.11 0.06 -0.08 -0.03 -0.01 0.00 -0.03 -0.06 0.01
φ8 0.00 0.00 0.00 0.00 0.00 -0.01 0.00 1.00 0.01 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
φ9 0.05 0.06 -0.02 -0.23 -0.05 -0.20 -0.18 0.01 1.00 -0.01 0.02 0.01 -0.05 0.12 0.00 -0.18 -0.04 -0.16
φ10 0.03 0.05 -0.01 -0.03 0.02 -0.56 -0.11 0.00 -0.01 1.00 -0.15 0.28 -0.09 0.09 0.00 -0.10 -0.15 -0.17
φ11 0.51 0.15 0.00 -0.09 -0.51 0.24 0.06 0.00 0.02 -0.15 1.00 -0.54 -0.07 0.08 0.00 -0.20 -0.01 -0.12
φ12 -0.10 0.06 -0.01 0.08 0.00 -0.12 -0.08 0.00 0.01 0.28 -0.54 1.00 -0.03 0.03 0.00 -0.04 -0.05 -0.06
φ13 -0.06 0.03 -0.03 0.00 0.11 -0.02 -0.03 0.00 -0.05 -0.09 -0.07 -0.03 1.00 0.12 0.00 0.37 0.42 0.37
φ14 0.03 0.00 -0.03 -0.13 -0.17 0.08 -0.01 0.00 0.12 0.09 0.08 0.03 0.12 1.00 0.00 -0.56 -0.17 -0.65
φ15 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 1.00 0.00 0.00 0.00
φ16 -0.07 -0.04 0.03 0.14 0.12 -0.30 -0.03 0.00 -0.18 -0.10 -0.20 -0.04 0.37 -0.56 0.00 1.00 0.42 0.86
φ17 -0.06 0.01 -0.01 0.03 0.02 0.04 -0.06 0.00 -0.04 -0.15 -0.01 -0.05 0.42 -0.17 0.00 0.42 1.00 0.34
φ18 -0.12 -0.04 0.02 0.14 0.09 -0.10 0.01 0.00 -0.16 -0.17 -0.12 -0.06 0.37 -0.65 0.00 0.86 0.34 1.00
* Model parameters, φi, were defined in Equations 7.25 through 7.31 in Section 7.4.

G 1
= a
G max γ 
1 +  
γ r 
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ 3 4

a = φ5
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin
 G max 
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )]

b = φ11 + φ12 * ln( N )

218
After concluding that the effects of geology (analyzed through geographic

location) and fines content (analyzed through soil groups) on the model

parameters were not very pronounced based on the analysis of subsets of the

database, all credible data (within the updated database) were analyzed as one

complete data set to calibrate the model. Calculation of mean normalized modulus

reduction and material damping curves and handling uncertainty associated with

these curves are discussed in the following chapters.

219
CHAPTER 9
PREDICTING NONLINEAR SOIL BEHAVIOR USING THE
CALIBRATED MODEL

9.1 INTRODUCTION

Proposed equations discussed in Section 7.4 have been calibrated using all

credible data in Section 8.3. The updated mean values of the model parameters

presented in Table 8.12 can be used to estimate normalized modulus reduction


and material damping curves for a broad range of soil types and loading

conditions.

Since the predictions are based on the model calibrated using all credible
data, the effects of a number of parameters regarding soil type (geology, fines

content, particle size, particle stiffness, etc.) are ignored in this model. The only

indicator of soil characteristics utilized in the estimation of nonlinear behavior is

plasticity index, PI.

In this chapter, estimation of nonlinear curves for a given soil plasticity

and loading condition is presented. Additionally, general trends based on these


estimated curves and their consistency with previous studies are discussed.

220
9.2 CALCULATION OF REFERENCE STRAIN, CURVATURE COEFFICIENT,
SMALL-STRAIN MATERIAL DAMPING RATIO AND THE SCALING
COEFFICIENT

The equations presented in Section 7.4.1 can be utilized to calculate

reference strain, curvature coefficient, small-strain material damping ratio and the

scaling coefficient by replacing parameters (φi) with their updated mean values
presented in Table 8.12 as follows:
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ
3 4
(9.1a)
a = φ5 (9.1b)
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )] (9.1c)

b = φ11 + φ12 * ln( N ) (9.1d)

where: σo’ = mean effective confining pressure (atm),


PI = soil plasticity (%),
OCR = overconsolidation ratio,

frq = loading frequency,

N = number of loading cycles,


φ1 = 0.0352,
φ2 = 0.0010,
φ3 = 0.3246,
φ4 = 0.3483,
φ5 = 0.9190,
φ6 = 0.8005,
φ7 = 0.0129,
φ8 = -0.1069,

221
φ9 = -0.2889,
φ10 = 0.2919,
φ11 = 0.6329, and
φ12 = -0.0057.
In this way, the relationship between the four model parameters (reference

strain, curvature coefficient, small-strain material damping ratio and the scaling

coefficient) discussed in Chapter Six, and soil plasticity and loading conditions
can be established based on statistical analysis of the database. These

relationships are also presented in a graphical form to assist the reader in


understanding the model characteristics and in utilizing the model.

Figure 9.1 shows a graphical tool that can be used to estimate reference

strain for given values of PI, OCR and in-situ mean effective stress. An example

solution is presented utilizing a clayey soil with PI = 60 % and moderate

overconsolidation (OCR = 4) subjected to 4 atm in-situ mean effective stress.


Starting with the PI and OCR of the soil, reference strain of the material is

estimated as if it were subjected to 1 atm confining pressure. This value is then

adjusted for the effect of confining pressure in the graphical solution.


Figure 9.2 shows the graphical solution for the scaling coefficient, b.

Calculation of the value for ten cycles of loading is presented as an example in

this figure.

222
σo' = 0.25 atm

σo' = 16 atm
σo' = 1 atm
σo' = 4 atm
0.30

OCR = 16
0.25
Reference Strain at 1 atm, %

0.20
OCR = 4
0.15

0.10
OCR = 1

0.05

0.00
100 80 60 40 20 0 0.01 0.1 1
Plasticity Index, % Reference Strain, %

Figure 9.1 Estimation of reference strain for given values of PI, OCR and in-
situ mean effective stress

0.64

0.63
Scaling Coefficient, b

0.62

0.61

0.60

0.59

0.58
1 10 100 1000
Number of Loading Cycles

Figure 9.2 Estimation of scaling coefficient for a given value of number of


loading cycles

223
Calculation of the small-strain material damping ratio is presented in

graphical form in Figure 9.3. As in the case of reference strain, small-strain

material damping ratio can be estimated graphically for given values of PI, OCR

in-situ mean effective stress and loading frequency. An example solution is

presented utilizing a clayey soil with PI = 60 % and moderate overconsolidation

(OCR = 4) subjected to 4 atm in-situ mean effective stress loaded at 10 Hz.

Starting with the PI and OCR of the soil, the small-strain material damping ratio

is estimated as if it were subjected to 1 atm confining pressure loaded at 1 Hz.

This value is adjusted for the effect of confining pressure and then for loading
frequency in the graphical solution.

9.3 ESTIMATION OF NORMALIZED MODULUS REDUCTION AND


MATERIAL DAMPING CURVES

Once the four model parameters (reference strain, curvature coefficient,

small-strain material damping ratio, and scaling coefficient) are calculated for the
soil plasticity and loading conditions, the equations presented in Chapter Six can

be utilized to estimate the normalized modulus reduction and material damping

curves as follows:

G 1
= a
(9.2a)
Gmax γ 
1 +  
γ r 
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin (9.2b)
 Gmax 

224
σo' = 0.25 atm
σo' = 16 atm
σo' = 4 atm
σo' = 1 atm
2.2
OCR = 1
2.0
Dmin at 1 atm and at 1 Hz, %

1.8 OCR = 4

1.6
OCR = 16
1.4

1.2

1.0

0.8
100 80 60 40 20 0 0.1 1 10
Plasticity Index, % Dmin at 1 Hz, %
10
f = 100 Hz
f = 10 Hz

f = 1 Hz
Dmin, %

0.1

Figure 9.3 Estimation of small-strain material damping ratio for given values of
PI, OCR, in-situ mean effective stress and loading frequency

225
G
where: = normalized shear modulus,
Gmax

γ = shearing strain (%),


γr = reference strain (%),
a = curvature coefficient,

Dmin = small-strain material damping ratio (%),


b = scaling coefficient,

DAdjusted = scaled and capped material damping (%),


DMa sin g = c1 DMa sin g ,a =1.0 + c 2 DMa sin g ,a =1.0 2 + c3 DMa sin g ,a =1.0 3 (%),
 γ +γ r  
 γ − γ r ln  
DMa sin g ,a =1.0 =
100 
4  γr  − 2 (%),
Π  γ2 
 
 γ +γr 

c1 = -1.1143a 2 + 1.8618a + 0.2523 ,


c 2 = 0.0805a 2 - 0.0710a - 0.0095 , and
c3 = - 0.0005a 2 + 0.0002a + 0.0003 .

Figure 9.4 shows the normalized modulus reduction and material damping

curves for the soil type and loading conditions presented in Section 9.2, a clayey

soil with PI = 60 % and moderate overconsolidation (OCR = 4) subjected to 4 atm


in-situ mean effective stress and ten cycles of loading at 10 Hz.

226
1.2
(a)
1.0

0.8

0.6
G/Gmax G/Gmax Prediction
0.4 γ r = 0.212 %
0.2 a = 0.92

0.0

25
Material Damping Prediction
20 γ r = 0.212 %
a = 0.92
15 Dmin = 1.65
D, %
b = 0.62
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 9.4 Estimated (a) normalized modulus reduction and (b) material
damping curves for the soil type and loading conditions discussed in
Section 9.2

227
It is important to note that the nonlinear behavior predicted by the model

is based on data collected over shearing strain amplitudes ranging from 1x10-5 %
to less than 1 %. As a result, extrapolation of the curves to higher strain

amplitudes is not recommended. Also, predicted material damping ratios at strain

amplitudes over 10 % will decrease to smaller values because of the damping


adjustment that introduces a cap on material damping. Consequently, the model

should never be utilized in modeling soil behavior at such high strain levels unless

the results are verified by additional tests performed at high strain amplitudes.

9.4 EFFECT OF OVERCONSOLIDATION RATIO, LOADING FREQUENCY AND


NUMBER OF LOADING CYCLES ON NONLINEAR SOIL BEHAVIOR

The effects of overconsolidation ratio, loading frequency and number of

loading cycles have been included in the model calibrated in Chapter Eight. The
results indicate that the effects of these variables on dynamic soil behavior are not

pronounced for the competent soils (that do not exhibit large volume change when

sheared at strains less than 1 %) investigated in this study.

Figure 9.5 presents the effect of overconsolidation ratio on nonlinear soil

behavior predicted by the calibrated model. Increasing overconsolidation ratio is

observed to result in a slight shift of the normalized modulus reduction and


material damping curves to higher strain amplitudes, along with a slight decrease

in small-strain material damping ratio. This effect is more pronounced for high
plasticity materials. These trends are consistent with those proposed by Hardin

and Drnevich (1972b).

228
1.2
(a)
1.0
0.8 G/Gmax Prediction
0.6 ( σo' = 1 atm, PI = 15 %,
G/Gmax N = 10 cycles, f = 1 Hz )
0.4 OCR = 1
0.2 OCR = 4
OCR = 16
0.0

25
Material Damping Prediction
20 ( σo' = 1 atm, PI = 15 %,
N = 10 cycles, f = 1 Hz )
15 OCR = 1
D, % OCR = 4
10 OCR = 16

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 9.5 Effect of overconsolidation ratio on (a) normalized modulus


reduction and (b) material damping curves predicted by the
calibrated model

229
In Figures 9.6 and 9.7, the effects of loading frequency and number of

loading cycles are shown. As discussed in Section 8.3, the model has been

formulated ignoring the effect of these two variables on the normalized modulus

reduction curve based on general trends observed during the course of this study.

Figure 9.6a and 9.7a are presented to clarify this issue.

Figure 9.6b presents the effect of loading frequency on the material

damping curve predicted by the calibrated model. An increase in small-strain

material damping ratio with increasing loading frequency is observed in this

figure. This effect is consistent with the trends reported in Stokoe et al. (1999).
In Figure 9.7b, the effect of number of loading cycles on the material

damping curve is presented. Increasing number of cycles results in a slight

decrease in the scaling coefficient causing a slight decrease of material damping

ratio at high strains. This general trend is also consistent with the trends reported

in Hardin and Drnevich (1972b) and Stokoe et al. (1999).

The database utilized to calibrate the four-parameter model consists of the

results of the first and tenth cycles of torsional shear tests performed at 1 Hz and
resonant column tests performed at the resonant frequency of the specimen that is

typically on the order of around 100 Hz. During resonant column testing, the
specimen is cycled about 1000 times. Thus, the combined effect of loading

frequency and number of loading cycles is presented in Figure 9.8 showing the

predicted difference between hypothetical test results.

230
1.2
(a)
1.0
0.8 G/Gmax Prediction
0.6 ( σo' = 1 atm, PI = 15 %,
G/Gmax N = 10 cycles, OCR = 1 )
0.4 f = 1 Hz
0.2 f = 10 Hz
f = 100 Hz
0.0

25
Material Damping Prediction
20 ( σo' = 1 atm, PI = 15 %,
N = 10 cycles, OCR = 1 )
15 f = 1 Hz
D, % f = 10 Hz
10 f = 100 Hz

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 9.6 Effect of loading frequency on (a) normalized modulus reduction


and (b) material damping curves predicted by the calibrated model

231
1.2
(a)
1.0
0.8 G/Gmax Prediction
0.6 ( σo' = 1 atm, PI = 15 %,
G/Gmax f = 1 Hz, OCR =1 )
0.4 N = 1 cycles
0.2 N = 10 cycles
N = 1000 cycles
0.0

25
Material Damping Prediction
20 ( σo' = 1 atm, PI = 15 %,
f = 1 Hz, OCR =1 )
15 N = 1 cycles
D, % N = 10 cycles
10 N = 1000 cycles

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 9.7 Effect of number of loading cycles on (a) normalized modulus


reduction and (b) material damping curves predicted by the
calibrated model

232
1.2
(a)
1.0
0.8
G/Gmax Prediction
0.6
G/Gmax ( σo' = 1 atm, PI = 15 %, OCR =1 )
0.4 f = 1 Hz, N = 1 cycles
f = 1 Hz, N = 10 cycles
0.2
f = 100 Hz, N = 1000 cycles
0.0

25
Material Damping Prediction
20 ( σo' = 1 atm, PI = 15 %, OCR =1 )
f = 1 Hz, N = 1 cycles
15 f = 1 Hz, N = 10 cycles
D, % f = 100 Hz, N = 1000 cycles
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 9.8 Comparison of (a) normalized modulus reduction and (b) material
damping curves predicted for resonant column and torsional shear
tests

233
9.5 EFFECT OF CONFINING PRESSURE ON NONLINEAR SOIL BEHAVIOR

The effect of confining pressure on normalized modulus reduction and

material damping curves predicted by the calibrated four-parameter model is

presented in Figure 9.9. The model shows the shift of normalized modulus

reduction and material damping curves to higher strain amplitudes with increasing
confining pressure along with a simultaneous decrease in small-strain material

damping ratio.
In Figure 9.10, the empirical curves proposed by Seed et al. (1986) are

presented. The comparison of the normalized modulus reduction curves predicted

by the calibrated model (Figure 9.9a) and the empirical curves proposed by Seed

et al. (1986) (Figure 9.10a) are presented in Figure 9.11a. The fact that the

nonlinear curves analyzed in Seed et al. (1986) were collected at low confining
pressures is supported by the close agreement between the mean Seed et al.

(1986) curve and the calibrated model curve for 1 atm. However, the comparison

of the material damping curves in Figure 9.11b shows that the material damping
values proposed by Seed et al. (1986) are higher than those encountered in the

course of this study. The discrepancy is believed to result from accuracy problems

in material damping measurements arising from the use of older generation cyclic

triaxial equipment employed in previous studies.

234
1.2
(a)
1.0
0.8
0.6
G/Gmax
0.4 G/Gmax Prediction
( PI = 0 %, N = 10 cycles,
0.2 f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 σo' = 0.25 atm
σo' = 1 atm
15 σo' = 4 atm
D, %
σo' = 16 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 9.9 Effect of confining pressure on (a) normalized modulus reduction


and (b) material damping curves predicted by the calibrated model

235
1.2
(a)

0.8

G/Gmax
0.4

0.0

20
Seed et al., (1986)
Average for Sands
15
Range

D, % 10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 9.10 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed for sands by Seed et al. (1986)

236
1.2
(a)
1.0
0.8
0.6 G/Gmax Prediction
G/Gmax ( PI = 0 %, N = 10 cycles,
0.4 f = 1 Hz, OCR = 1 )
0.2 Seed et al. (1986)
0.0

25
Material Damping Prediction
20 σo' = 0.25 atm
σo' = 1 atm
15 σo' = 4 atm
D, %
σo' = 16 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 9.11 Comparison of the effect of confining pressure on nonlinear soil


behavior of sand (PI = 0 %) predicted by the calibrated model and
empirical curves proposed for sands by Seed et al. (1986)

237
9.6 EFFECT OF SOIL TYPE ON NONLINEAR SOIL BEHAVIOR

The effect of soil plasticity on normalized modulus reduction and material

damping curves predicted by the calibrated four-parameter model is presented in

Figure 9.12. The model shows shifts in the normalized modulus reduction and

material damping curves to higher strain amplitudes with increasing soil plasticity
along with a simultaneous increase in the small-strain material damping ratio.

In Figure 9.13, the empirical curves proposed by Vucetic and Dobry


(1991) are presented. Comparison of the normalized modulus reduction curves

predicted by the calibrated model and the empirical curves proposed by Vucetic

and Dobry (1991) is presented in Figure 9.14a. As seen in the figure, the general

trend presented by Vucetic and Dobry (1991) agrees with this work. However, the

effect of soil plasticity presented by Vucetic and Dobry (1991) is more


pronounced than observed in this study.

Comparison of the material damping curves predicted by the calibrated

model and the empirical curves proposed by Vucetic and Dobry (1991) is
presented in Figure 9.14b. As seen in the figure, the material damping curves

proposed by Vucetic and Dobry (1991) also indicate a more pronounced effect of

soil plasticity. Also, as discussed in Section 5.3, the Vucetic and Dobry (1991)

damping curves do not accurately model the observed increase in small-strain

material damping ratio with increasing soil plasticity. As in the case of Seed et al.

(1986), the discrepancy is believed to be a result of accuracy problems in damping


measurements arising from the use of older generation equipment in previous

studies.

238
1.2
(a)
1.0
0.8
0.6
G/Gmax
G/Gmax Prediction
0.4
( σo' = 1 atm, N = 10 cycles,
0.2 f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 PI = 0 %
PI = 15 %
15 PI = 30 %
D, % PI = 50 %
10 PI = 100 %

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 9.12 Effect of soil plasticity on (a) normalized modulus reduction and (b)
material damping curves predicted by the calibrated model

239
1.2
(a)

0.8

G/Gmax
0.4

0.0

20
Vucetic and Dobry (1991)
15 Non-Plastic
PI = 15 %
D, % 10 PI = 30 %
PI = 50 %
PI = 100 %
5
PI = 200 %
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ, %

Figure 9.13 Empirical (a) normalized modulus reduction, and (b) material
damping curves proposed by Vucetic and Dobry (1991)

240
1.2
(a)
1.0
0.8
G/Gmax Prediction
0.6
G/Gmax ( σo' = 1 atm, N = 10 cycles,
0.4 f = 1 Hz, OCR = 1 )

0.2 Vucetic and Dobry (1991)


0.0

25
Material Damping Prediction
20 PI = 0 %
PI = 15 %
15 PI = 30 %
D, % PI = 50 %
10 PI = 100 %

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ,%

Figure 9.14 Comparison of the effect of soil plasticity on nonlinear soil behavior
predicted by the calibrated model and empirical curves proposed by
Vucetic and Dobry (1991)

241
9.7 EFFECTS OF CONFINING PRESSURE AND SOIL TYPE ON STRESS-
STRAIN CURVES

The effects of confining pressure and soil type on normalized modulus

reduction curves predicted by the calibrated model have been evaluated in

Sections 9.5 and 9.6. Since the normalized modulus reduction curves analyzed as

part of this study are actually secant shear moduli scaled down using small-strain
values, stress-strain curves can be evaluated using the calibrated model.

The relationship between shear stress and secant shear modulus is:
τ = G *γ (9.3)

where; τ = shear stress (MPa),


γ = shearing strain,
 G 
G = Gmax*   = shear modulus (MPa),
 Gmax 
G 1
= a
,
Gmax γ 
1 +  
γ r 

Gmax = ρ*Vs2 = small-strain shear modulus (MPa),


ρ = mass density (kg/m3), and
Vs = shear wave velocity (m/sec).

242
Equation 9.3 illustrates that, in order to estimate stress-strain curves based on the

four-parameter model, the impact of confining pressure and soil type on shear

wave velocity has to be evaluated. The relationship between these parameters and

in-situ shear wave velocity measurements can be assessed utilizing the same

database that was used to calibrate the four-parameter model. Eighty seven of the

specimens within the database were sampled from sites where in-situ shear wave

velocity measurements had been performed (Chiara, 2001). A least-squares fit to

part of the data suggests the following relationship:


Vs = 300 * σ o ' 0.27 −2.44 * PI (9.4)

where; σo’ = mean effective confining pressure (atm), and


PI = plasticity index (%).
The comparison of the predicted values of shear wave velocity and

measured values is presented in Figure 9.15 so that the reader can evaluate the

quality of the fit. Equation 9.4 is not part of the calibrated model recommended

for evaluating dynamic soil behavior and is utilized only as a starting point to

generate stress-strain curves presented in this section.


The effects of confining pressure and soil type on stress-strain curves

predicted using the calibrated four-parameter model are presented in Figures 9.16
and 9.17, respectively. Figures 9.16a and 9.17a show the predictions for shearing

strains up to 1 %. Figures 9.16b and 9.17b show the predictions for shearing

strains up to 0.01 % so that the characteristics of the curves at smaller strains can

be presented to the reader.

243
1000

Predicted Shear Wave Velocity, m/sec


800

600

400

200

0
0 200 400 600 800 1000

Measured Shear Wave Velocity, m/sec

Figure 9.15 Comparison of the measured in-situ shear wave velocities and values
predicted using Equation 9.4

The model successfully predicts the increase in shear strength with

increasing confining pressure as shown in Figure 9.16. Figure 9.17 shows a

comparison of predicted stress-strain curves for soils with different plasticities.

Figure 9.18 shows part of the data in Figure 9.17 in an effort to compare the

behavior of a sand (PI = 0 %) and a moderate plasticity clay (PI = 30 %). It is

important to note the similarity between this figure and Figure 5.1b from Hardin
and Drnevich (1972b).

244
300
σo' = 0.25 atm (a)
σo' = 1 atm
250
σo' = 4 atm

200

τ, kPa 150

100

50

0
0.0 0.2 0.4 0.6 0.8 1.0

Shearing Strain, γ,%


10
(b)

6
τ, kPa

σo' = 0.25 atm


2 σo' = 1 atm
σo' = 4 atm
0
0.000 0.002 0.004 0.006 0.008 0.010

Shearing Strain, γ,%

Figure 9.16 Effect of confining pressure on stress-strain curve predicted by the


calibrated model for shearing strains ranging (a) from γ = 0 to 1 %
and (b) from γ = 0 to 0.01 %

245
100
PI = 0 % (a)
PI = 15 %
80 PI = 30 %
PI = 50 %

60
τ, kPa
40

20

0
0.0 0.2 0.4 0.6 0.8 1.0

Shearing Strain, γ,%


10
(b)

6
τ, kPa
4
PI = 0 %
PI = 15 %
2
PI = 30 %
PI = 50 %
0
0.000 0.002 0.004 0.006 0.008 0.010

Shearing Strain, γ,%

Figure 9.17 Effect of soil plasticity on stress-strain curve predicted by the


calibrated model for shearing strains ranging (a) from γ = 0 to 1 %
and (b) from γ = 0 to 0.01 %

246
100
PI = 0 % (a)
PI = 30 %
80

60
τ, kPa
40

20

0
0.0 0.2 0.4 0.6 0.8 1.0

Shearing Strain, γ,%


10
(b)

6
τ, kPa
4

2
PI = 0 %
PI = 30 %
0
0.000 0.002 0.004 0.006 0.008 0.010

Shearing Strain, γ,%

Figure 9.18 Comparison of the stress-strain curves of a sand and a moderate


plasticity clay based on the calibrated model for shearing strains
ranging (a) from γ = 0 to 1 % and (b) from γ = 0 to 0.01 %

247
9.8 SUMMARY

In this chapter, equations for the calibrated model are presented along with

graphical solutions that can be utilized in predicting normalized modulus

reduction and material damping curves for a given soil type and a given loading

condition.
The general trends of the predicted curves are briefly discussed. The

results indicate that soil plasticity and mean effective confining pressure are the
two most important parameters that control nonlinear behavior of “competent”

soils strained up to γ = 1 %. The comparison of the predicted curves with those


presented in the literature indicates a general agreement with the trends proposed
by other researchers. However, this comparison also highlights discrepancies in

material damping measurements resulting from the limitations of older generation

testing equipment utilized in previous studies.

Since stress-strain curves are related to normalized shear modulus curves,

a brief discussion regarding the predicted stress-strain curves based on the


calibrated model is also presented herein in an effort to bridge the gap between

traditional geotechnical engineering and soil dynamics. The findings indicate that
the results of dynamic tests can also be utilized in the traditional geotechnical

engineering applications that require modeling soil behavior at working strains.

248
CHAPTER 10
RECOMMENDED NORMALIZED MODULUS REDUCTION
AND MATERIAL DAMPING CURVES

10.1 INTRODUCTION

Mean values of the normalized shear modulus and the material damping

ratio (predicted by the calibrated model) at strain amplitudes ranging from 1x10-5
% to 1 % are presented in this chapter. As discussed in Chapter Nine, the mean

values of model parameters can be utilized to construct normalized modulus


reduction and material damping curves for different soil types and loading

conditions. However, the reader must use caution when a soil type or loading

condition not represented in the database is to be evaluated with these equations.


Since the impact of overconsolidation ratio is relatively small and ten

cycles at 1 Hz loading frequency closely represents the characteristics of

earthquake shaking, these parameters are fixed for the recommended curves. In

this chapter, recommended normalized modulus reduction and material damping

curves are presented for soils with a broad range of plasticities confined at a broad

range of mean effective stresses.


These curves are presented from two different perspectives so that the

reader can interpolate the data for different values of soil plasticity and confining
pressure. If the reader has to extrapolate for soil plasticities and confining

pressures not represented in the database, use of caution is suggested.

249
10.2 EFFECT OF PI AT A GIVEN MEAN EFFECTIVE STRESS

Figures 10.1 through 10.4 show the effect of PI on nonlinear soil behavior

at 0.25, 1.0, 4.0 and 16 atm, respectively. These normalized modulus and material

damping curves are presented so that the reader can interpolate these relationships

for soils with different plasticities. Also, these curves are tabulated in Tables 10.1
through 10.8. The figures and tables are organized so that the G/Gmax – log γ and
D – log γ curves are followed on the next page by the associated tables.

10.3 EFFECT OF MEAN EFFECTIVE STRESS ON A SOIL WITH GIVEN


PLASTICITY

Figures 10.5 through 10.9 show the effect of mean effective stress on

nonlinear behavior of soils with 0, 15, 30, 50 and 100 % plasticity, respectively.
These normalized modulus and material damping curves are presented so that the

reader can interpolate these relationships for soil layers at different depths

confined under different mean effective stresses. Also, these curves are tabulated

in Tables 10.9 through 10.18. The figures and tables are organized so that the

G/Gmax – log γ and D – log γ curves are followed on the next page by the
associated tables.

10.4 IMPACT OF UTILIZING THE RECOMMENDED CURVES ON


EARTHQUAKE RESPONSE PREDICTIONS OF DEEP SOIL SITES

The impact of utilizing the recommended curves when assigning nonlinear


soil properties in site response analyses of deep (>50 m) soil sites is discussed in

this section. This point is addressed because site response analyses are often
performed using average, pressure-independent generic curves.

250
1.2
(a)
1.0

0.8

0.6
G/Gmax
G/Gmax Prediction
0.4
( σo' = 0.25 atm, N = 10 cycles,
0.2 f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 PI = 0 %
PI = 15 %
15 PI = 30 %
D, % PI = 50 %
10 PI = 100 %

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 10.1 Effect of PI on (a) normalized modulus reduction and (b) material
damping curves at 0.25 atm confining pressure

251
Table 10.1 Effect of PI on normalized modulus reduction curve: σo’ = 0.25 atm

Shearing Strain (%) PI = 0 % PI = 15 % PI = 30 % PI = 50 % PI = 100 %


1.00E-05 0.999 0.999 1.000 1.000 1.000
2.20E-05 0.998 0.999 0.999 0.999 1.000
4.84E-05 0.996 0.997 0.998 0.998 0.999
1.00E-04 0.993 0.995 0.996 0.997 0.998
2.20E-04 0.986 0.990 0.992 0.994 0.996
4.84E-04 0.971 0.979 0.983 0.987 0.991
1.00E-03 0.944 0.959 0.968 0.975 0.983
2.20E-03 0.891 0.919 0.936 0.949 0.966
4.84E-03 0.799 0.847 0.876 0.900 0.932
1.00E-02 0.671 0.739 0.783 0.822 0.876
2.20E-02 0.497 0.579 0.637 0.692 0.774
4.84E-02 0.324 0.400 0.459 0.521 0.625
1.00E-01 0.197 0.255 0.303 0.358 0.461
2.20E-01 0.107 0.142 0.174 0.213 0.293
4.84E-01 0.055 0.074 0.093 0.116 0.167
1.00E+00 0.029 0.040 0.050 0.063 0.093

Table 10.2 Effect of PI on material damping curve: σo’ = 0.25 atm

Shearing Strain (%) PI = 0 % PI = 15 % PI = 30 % PI = 50 % PI = 100 %


1.00E-05 1.201 1.489 1.778 2.164 3.129
2.20E-05 1.207 1.493 1.781 2.166 3.131
4.84E-05 1.226 1.506 1.791 2.174 3.136
1.00E-04 1.257 1.528 1.808 2.187 3.144
2.20E-04 1.330 1.579 1.848 2.217 3.163
4.84E-04 1.487 1.690 1.933 2.282 3.204
1.00E-03 1.792 1.906 2.101 2.411 3.286
2.20E-03 2.458 2.387 2.476 2.702 3.472
4.84E-03 3.762 3.358 3.249 3.310 3.868
1.00E-02 5.821 4.977 4.581 4.386 4.593
2.20E-02 9.097 7.778 7.010 6.441 6.070
4.84E-02 12.993 11.489 10.477 9.589 8.579
1.00E-01 16.376 15.064 14.088 13.137 11.798
2.20E-01 19.181 18.334 17.640 16.904 15.716
4.84E-01 20.829 20.515 20.208 19.849 19.213
1.00E+00 21.393 21.507 21.542 21.547 21.544

252
1.2
(a)
1.0

0.8

0.6
G/Gmax
G/Gmax Prediction
0.4
( σo' = 1 atm, N = 10 cycles,
0.2 f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 PI = 0 %
PI = 15 %
15 PI = 30 %
D, % PI = 50 %
10 PI = 100 %

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 10.2 Effect of PI on (a) normalized modulus reduction and (b) material
damping curves at 1.0 atm confining pressure

253
Table 10.3 Effect of PI on normalized modulus reduction curve: σo’ = 1.0 atm

Shearing Strain (%) PI = 0 % PI = 15 % PI = 30 % PI = 50 % PI = 100 %


1.00E-05 0.999 1.000 1.000 1.000 1.000
2.20E-05 0.999 0.999 0.999 1.000 1.000
4.84E-05 0.998 0.998 0.999 0.999 0.999
1.00E-04 0.995 0.997 0.997 0.998 0.999
2.20E-04 0.991 0.993 0.995 0.996 0.997
4.84E-04 0.981 0.986 0.989 0.992 0.994
1.00E-03 0.964 0.973 0.979 0.984 0.989
2.20E-03 0.928 0.947 0.958 0.967 0.978
4.84E-03 0.861 0.896 0.917 0.934 0.956
1.00E-02 0.761 0.816 0.849 0.878 0.917
2.20E-02 0.607 0.682 0.732 0.778 0.843
4.84E-02 0.428 0.509 0.569 0.629 0.722
1.00E-01 0.277 0.348 0.404 0.465 0.571
2.20E-01 0.157 0.205 0.248 0.296 0.392
4.84E-01 0.083 0.111 0.137 0.169 0.238
1.00E+00 0.044 0.060 0.076 0.095 0.138

Table 10.4 Effect of PI on material damping curve: σo’ = 1.0 atm

Shearing Strain (%) PI = 0 % PI = 15 % PI = 30 % PI = 50 % PI = 100 %


1.00E-05 0.804 0.997 1.191 1.450 2.096
2.20E-05 0.808 1.000 1.193 1.451 2.097
4.84E-05 0.820 1.008 1.199 1.456 2.100
1.00E-04 0.839 1.021 1.209 1.464 2.105
2.20E-04 0.884 1.053 1.234 1.482 2.117
4.84E-04 0.982 1.122 1.287 1.523 2.143
1.00E-03 1.174 1.257 1.392 1.603 2.193
2.20E-03 1.602 1.562 1.628 1.786 2.309
4.84E-03 2.474 2.198 2.128 2.175 2.560
1.00E-02 3.953 3.317 3.028 2.888 3.029
2.20E-02 6.579 5.440 4.803 4.343 4.029
4.84E-02 10.184 8.650 7.664 6.824 5.876
1.00E-01 13.788 12.217 11.092 10.024 8.541
2.20E-01 17.199 15.951 14.966 13.941 12.279
4.84E-01 19.565 18.829 18.185 17.458 16.132
1.00E+00 20.716 20.460 20.178 19.815 19.069

254
1.2
(a)
1.0

0.8

0.6
G/Gmax
G/Gmax Prediction
0.4
( σo' = 4 atm, N = 10 cycles,
0.2 f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 PI = 0 %
PI = 15 %
15 PI = 30 %
D, % PI = 50 %
10 PI = 100 %

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 10.3 Effect of PI on (a) normalized modulus reduction and (b) material
damping curves at 4.0 atm confining pressure

255
Table 10.5 Effect of PI on normalized modulus reduction curve: σo’ = 4.0 atm

Shearing Strain (%) PI = 0 % PI = 15 % PI = 30 % PI = 50 % PI = 100 %


1.00E-05 1.000 1.000 1.000 1.000 1.000
2.20E-05 0.999 1.000 1.000 1.000 1.000
4.84E-05 0.998 0.999 0.999 0.999 1.000
1.00E-04 0.997 0.998 0.998 0.999 0.999
2.20E-04 0.994 0.996 0.997 0.997 0.998
4.84E-04 0.988 0.991 0.993 0.995 0.996
1.00E-03 0.976 0.983 0.986 0.989 0.993
2.20E-03 0.952 0.965 0.972 0.978 0.986
4.84E-03 0.906 0.931 0.945 0.956 0.971
1.00E-02 0.832 0.873 0.898 0.918 0.945
2.20E-02 0.706 0.770 0.810 0.845 0.893
4.84E-02 0.538 0.618 0.673 0.725 0.802
1.00E-01 0.374 0.454 0.514 0.575 0.675
2.20E-01 0.225 0.287 0.339 0.396 0.501
4.84E-01 0.123 0.163 0.199 0.241 0.327
1.00E+00 0.067 0.091 0.113 0.140 0.200

Table 10.6 Effect of PI on material damping curve: σo’ = 4.0 atm

Shearing Strain (%) PI = 0 % PI = 15 % PI = 30 % PI = 50 % PI = 100 %


1.00E-05 0.539 0.668 0.798 0.971 1.404
2.20E-05 0.541 0.670 0.799 0.972 1.405
4.84E-05 0.548 0.675 0.803 0.975 1.407
1.00E-04 0.560 0.683 0.809 0.980 1.410
2.20E-04 0.588 0.703 0.824 0.991 1.417
4.84E-04 0.649 0.745 0.857 1.016 1.433
1.00E-03 0.769 0.829 0.922 1.066 1.464
2.20E-03 1.039 1.021 1.070 1.180 1.537
4.84E-03 1.607 1.428 1.388 1.426 1.693
1.00E-02 2.618 2.173 1.977 1.886 1.991
2.20E-02 4.572 3.684 3.206 2.871 2.648
4.84E-02 7.621 6.235 5.387 4.693 3.934
1.00E-01 11.134 9.482 8.357 7.333 5.972
2.20E-01 14.946 13.400 12.231 11.056 9.226
4.84E-01 17.990 16.866 15.935 14.917 13.118
1.00E+00 19.792 19.158 18.571 17.876 16.513

256
1.2
(a)
1.0

0.8

0.6
G/Gmax
G/Gmax Prediction
0.4
( σo' = 16 atm, N = 10 cycles,
0.2 f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 PI = 0 %
PI = 15 %
15 PI = 30 %
D, % PI = 50 %
10 PI = 100 %

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 10.4 Effect of PI on (a) normalized modulus reduction and (b) material
damping curves at 16 atm confining pressure

257
Table 10.7 Effect of PI on normalized modulus reduction curve: σo’ = 16 atm

Shearing Strain (%) PI = 0 % PI = 15 % PI = 30 % PI = 50 % PI = 100 %


1.00E-05 1.000 1.000 1.000 1.000 1.000
2.20E-05 1.000 1.000 1.000 1.000 1.000
4.84E-05 0.999 0.999 0.999 1.000 1.000
1.00E-04 0.998 0.999 0.999 0.999 0.999
2.20E-04 0.996 0.997 0.998 0.998 0.999
4.84E-04 0.992 0.994 0.996 0.997 0.998
1.00E-03 0.985 0.989 0.991 0.993 0.996
2.20E-03 0.969 0.977 0.982 0.986 0.991
4.84E-03 0.938 0.954 0.964 0.972 0.981
1.00E-02 0.885 0.915 0.932 0.946 0.964
2.20E-02 0.789 0.839 0.869 0.895 0.929
4.84E-02 0.645 0.716 0.763 0.804 0.863
1.00E-01 0.482 0.564 0.623 0.679 0.764
2.20E-01 0.311 0.386 0.444 0.506 0.610
4.84E-01 0.179 0.233 0.279 0.331 0.431
1.00E+00 0.101 0.135 0.166 0.203 0.280

Table 10.8 Effect of PI on material damping curve: σo’ = 16 atm

Shearing Strain (%) PI = 0 % PI = 15 % PI = 30 % PI = 50 % PI = 100 %


1.00E-05 0.361 0.448 0.534 0.650 0.941
2.20E-05 0.362 0.449 0.535 0.651 0.941
4.84E-05 0.367 0.452 0.538 0.653 0.942
1.00E-04 0.374 0.457 0.541 0.656 0.944
2.20E-04 0.391 0.469 0.551 0.663 0.949
4.84E-04 0.429 0.495 0.571 0.678 0.958
1.00E-03 0.503 0.547 0.611 0.709 0.978
2.20E-03 0.673 0.667 0.704 0.780 1.023
4.84E-03 1.035 0.924 0.903 0.934 1.120
1.00E-02 1.702 1.407 1.281 1.227 1.308
2.20E-02 3.075 2.433 2.100 1.871 1.729
4.84E-02 5.449 4.318 3.659 3.138 2.589
1.00E-01 8.573 7.021 6.022 5.151 4.049
2.20E-01 12.483 10.780 9.557 8.381 6.651
4.84E-01 16.070 14.619 13.472 12.268 10.241
1.00E+00 18.528 17.522 16.655 15.677 13.847

258
1.2
(a)
1.0
0.8
0.6
G/Gmax
0.4 G/Gmax Prediction
( PI = 0 %, N = 10 cycles,
0.2 f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 σo' = 0.25 atm
σo' = 1 atm
15 σo' = 4 atm
D, %
σo' = 16 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 10.5 Effect of mean effective stress on (a) normalized modulus reduction
and (b) material damping curves of a nonplastic soil

259
Table 10.9 Effect of σo’ on normalized modulus reduction curve: PI = 0 %

Shearing Strain (%) σo' = 0.25 atm σo' = 1.0 atm σo' = 4.0 atm σo' = 16 atm
1.00E-05 0.999 0.999 1.000 1.000
2.20E-05 0.998 0.999 0.999 1.000
4.84E-05 0.996 0.998 0.998 0.999
1.00E-04 0.993 0.995 0.997 0.998
2.20E-04 0.986 0.991 0.994 0.996
4.84E-04 0.971 0.981 0.988 0.992
1.00E-03 0.944 0.964 0.976 0.985
2.20E-03 0.891 0.928 0.952 0.969
4.84E-03 0.799 0.861 0.906 0.938
1.00E-02 0.671 0.761 0.832 0.885
2.20E-02 0.497 0.607 0.706 0.789
4.84E-02 0.324 0.428 0.538 0.645
1.00E-01 0.197 0.277 0.374 0.482
2.20E-01 0.107 0.157 0.225 0.311
4.84E-01 0.055 0.083 0.123 0.179
1.00E+00 0.029 0.044 0.067 0.101

Table 10.10 Effect of σo’ on material damping curve: PI = 0 %

Shearing Strain (%) σo' = 0.25 atm σo' = 1.0 atm σo' = 4.0 atm σo' = 16 atm
1.00E-05 1.201 0.804 0.539 0.361
2.20E-05 1.207 0.808 0.541 0.362
4.84E-05 1.226 0.820 0.548 0.367
1.00E-04 1.257 0.839 0.560 0.374
2.20E-04 1.330 0.884 0.588 0.391
4.84E-04 1.487 0.982 0.649 0.429
1.00E-03 1.792 1.174 0.769 0.503
2.20E-03 2.458 1.602 1.039 0.673
4.84E-03 3.762 2.474 1.607 1.035
1.00E-02 5.821 3.953 2.618 1.702
2.20E-02 9.097 6.579 4.572 3.075
4.84E-02 12.993 10.184 7.621 5.449
1.00E-01 16.376 13.788 11.134 8.573
2.20E-01 19.181 17.199 14.946 12.483
4.84E-01 20.829 19.565 17.990 16.070
1.00E+00 21.393 20.716 19.792 18.528

260
1.2
(a)
1.0
0.8
0.6
G/Gmax
0.4 G/Gmax Prediction
( PI = 15 %, N = 10 cycles,
0.2 f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 σo' = 0.25 atm
σo' = 1 atm
15 σo' = 4 atm
D, %
σo' = 16 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 10.6 Effect of mean effective stress on (a) normalized modulus reduction
and (b) material damping curves of a soil with PI = 15 %

261
Table 10.11 Effect of σo’ on normalized modulus reduction curve: PI = 15 %

Shearing Strain (%) σo' = 0.25 atm σo' = 1.0 atm σo' = 4.0 atm σo' = 16 atm
1.00E-05 0.999 1.000 1.000 1.000
2.20E-05 0.999 0.999 1.000 1.000
4.84E-05 0.997 0.998 0.999 0.999
1.00E-04 0.995 0.997 0.998 0.999
2.20E-04 0.990 0.993 0.996 0.997
4.84E-04 0.979 0.986 0.991 0.994
1.00E-03 0.959 0.973 0.983 0.989
2.20E-03 0.919 0.947 0.965 0.977
4.84E-03 0.847 0.896 0.931 0.954
1.00E-02 0.739 0.816 0.873 0.915
2.20E-02 0.579 0.682 0.770 0.839
4.84E-02 0.400 0.509 0.618 0.716
1.00E-01 0.255 0.348 0.454 0.564
2.20E-01 0.142 0.205 0.287 0.386
4.84E-01 0.074 0.111 0.163 0.233
1.00E+00 0.040 0.060 0.091 0.135

Table 10.12 Effect of σo’ on material damping curve: PI = 15 %

Shearing Strain (%) σo' = 0.25 atm σo' = 1.0 atm σo' = 4.0 atm σo' = 16 atm
1.00E-05 1.489 0.997 0.668 0.448
2.20E-05 1.493 1.000 0.670 0.449
4.84E-05 1.506 1.008 0.675 0.452
1.00E-04 1.528 1.021 0.683 0.457
2.20E-04 1.579 1.053 0.703 0.469
4.84E-04 1.690 1.122 0.745 0.495
1.00E-03 1.906 1.257 0.829 0.547
2.20E-03 2.387 1.562 1.021 0.667
4.84E-03 3.358 2.198 1.428 0.924
1.00E-02 4.977 3.317 2.173 1.407
2.20E-02 7.778 5.440 3.684 2.433
4.84E-02 11.489 8.650 6.235 4.318
1.00E-01 15.064 12.217 9.482 7.021
2.20E-01 18.334 15.951 13.400 10.780
4.84E-01 20.515 18.829 16.866 14.619
1.00E+00 21.507 20.460 19.158 17.522

262
1.2
(a)
1.0
0.8
0.6
G/Gmax
0.4 G/Gmax Prediction
( PI = 30 %, N = 10 cycles,
0.2
f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 σo' = 0.25 atm
σo' = 1 atm
15 σo' = 4 atm
D, %
σo' = 16 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 10.7 Effect of mean effective stress on (a) normalized modulus reduction
and (b) material damping curves of a soil with PI = 30 %

263
Table 10.13 Effect of σo’ on normalized modulus reduction curve: PI = 30 %

Shearing Strain (%) σo' = 0.25 atm σo' = 1.0 atm σo' = 4.0 atm σo' = 16 atm
1.00E-05 1.000 1.000 1.000 1.000
2.20E-05 0.999 0.999 1.000 1.000
4.84E-05 0.998 0.999 0.999 0.999
1.00E-04 0.996 0.997 0.998 0.999
2.20E-04 0.992 0.995 0.997 0.998
4.84E-04 0.983 0.989 0.993 0.996
1.00E-03 0.968 0.979 0.986 0.991
2.20E-03 0.936 0.958 0.972 0.982
4.84E-03 0.876 0.917 0.945 0.964
1.00E-02 0.783 0.849 0.898 0.932
2.20E-02 0.637 0.732 0.810 0.869
4.84E-02 0.459 0.569 0.673 0.763
1.00E-01 0.303 0.404 0.514 0.623
2.20E-01 0.174 0.248 0.339 0.444
4.84E-01 0.093 0.137 0.199 0.279
1.00E+00 0.050 0.076 0.113 0.166

Table 10.14 Effect of σo’ on material damping curve: PI = 30 %

Shearing Strain (%) σo' = 0.25 atm σo' = 1.0 atm σo' = 4.0 atm σo' = 16 atm
1.00E-05 1.778 1.191 0.798 0.534
2.20E-05 1.781 1.193 0.799 0.535
4.84E-05 1.791 1.199 0.803 0.538
1.00E-04 1.808 1.209 0.809 0.541
2.20E-04 1.848 1.234 0.824 0.551
4.84E-04 1.933 1.287 0.857 0.571
1.00E-03 2.101 1.392 0.922 0.611
2.20E-03 2.476 1.628 1.070 0.704
4.84E-03 3.249 2.128 1.388 0.903
1.00E-02 4.581 3.028 1.977 1.281
2.20E-02 7.010 4.803 3.206 2.100
4.84E-02 10.477 7.664 5.387 3.659
1.00E-01 14.088 11.092 8.357 6.022
2.20E-01 17.640 14.966 12.231 9.557
4.84E-01 20.208 18.185 15.935 13.472
1.00E+00 21.542 20.178 18.571 16.655

264
1.2
(a)
1.0
0.8
0.6
G/Gmax
0.4 G/Gmax Prediction
( PI = 50 %, N = 10 cycles,
0.2 f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 σo' = 0.25 atm
σo' = 1 atm
15 σo' = 4 atm
D, %
σo' = 16 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 10.8 Effect of mean effective stress on (a) normalized modulus reduction
and (b) material damping curves of a soil with PI = 50 %

265
Table 10.15 Effect of σo’ on normalized modulus reduction curve: PI = 50 %

Shearing Strain (%) σo' = 0.25 atm σo' = 1.0 atm σo' = 4.0 atm σo' = 16 atm
1.00E-05 1.000 1.000 1.000 1.000
2.20E-05 0.999 1.000 1.000 1.000
4.84E-05 0.998 0.999 0.999 1.000
1.00E-04 0.997 0.998 0.999 0.999
2.20E-04 0.994 0.996 0.997 0.998
4.84E-04 0.987 0.992 0.995 0.997
1.00E-03 0.975 0.984 0.989 0.993
2.20E-03 0.949 0.967 0.978 0.986
4.84E-03 0.900 0.934 0.956 0.972
1.00E-02 0.822 0.878 0.918 0.946
2.20E-02 0.692 0.778 0.845 0.895
4.84E-02 0.521 0.629 0.725 0.804
1.00E-01 0.358 0.465 0.575 0.679
2.20E-01 0.213 0.296 0.396 0.506
4.84E-01 0.116 0.169 0.241 0.331
1.00E+00 0.063 0.095 0.140 0.203

Table 10.16 Effect of σo’ on material damping curve: PI = 50 %

Shearing Strain (%) σo' = 0.25 atm σo' = 1.0 atm σo' = 4.0 atm σo' = 16 atm
1.00E-05 2.164 1.450 0.971 0.650
2.20E-05 2.166 1.451 0.972 0.651
4.84E-05 2.174 1.456 0.975 0.653
1.00E-04 2.187 1.464 0.980 0.656
2.20E-04 2.217 1.482 0.991 0.663
4.84E-04 2.282 1.523 1.016 0.678
1.00E-03 2.411 1.603 1.066 0.709
2.20E-03 2.702 1.786 1.180 0.780
4.84E-03 3.310 2.175 1.426 0.934
1.00E-02 4.386 2.888 1.886 1.227
2.20E-02 6.441 4.343 2.871 1.871
4.84E-02 9.589 6.824 4.693 3.138
1.00E-01 13.137 10.024 7.333 5.151
2.20E-01 16.904 13.941 11.056 8.381
4.84E-01 19.849 17.458 14.917 12.268
1.00E+00 21.547 19.815 17.876 15.677

266
1.2
(a)
1.0
0.8
0.6
G/Gmax
0.4 G/Gmax Prediction
( PI = 100 %, N = 10 cycles,
0.2 f = 1 Hz, OCR = 1 )
0.0

25
Material Damping Prediction
20 σo' = 0.25 atm
σo' = 1 atm
15 σo' = 4 atm
D, %
σo' = 16 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 10.9 Effect of mean effective stress on (a) normalized modulus reduction
and (b) material damping curves of a soil with PI = 100 %

267
Table 10.17 Effect of σo’ on normalized modulus reduction curve: PI = 100 %

Shearing Strain (%) σo' = 0.25 atm σo' = 1.0 atm σo' = 4.0 atm σo' = 16 atm
1.00E-05 1.000 1.000 1.000 1.000
2.20E-05 1.000 1.000 1.000 1.000
4.84E-05 0.999 0.999 1.000 1.000
1.00E-04 0.998 0.999 0.999 0.999
2.20E-04 0.996 0.997 0.998 0.999
4.84E-04 0.991 0.994 0.996 0.998
1.00E-03 0.983 0.989 0.993 0.996
2.20E-03 0.966 0.978 0.986 0.991
4.84E-03 0.932 0.956 0.971 0.981
1.00E-02 0.876 0.917 0.945 0.964
2.20E-02 0.774 0.843 0.893 0.929
4.84E-02 0.625 0.722 0.802 0.863
1.00E-01 0.461 0.571 0.675 0.764
2.20E-01 0.293 0.392 0.501 0.610
4.84E-01 0.167 0.238 0.327 0.431
1.00E+00 0.093 0.138 0.200 0.280

Table 10.18 Effect of σo’ on material damping curve: PI = 100 %

Shearing Strain (%) σo' = 0.25 atm σo' = 1.0 atm σo' = 4.0 atm σo' = 16 atm
1.00E-05 3.129 2.096 1.404 0.941
2.20E-05 3.131 2.097 1.405 0.941
4.84E-05 3.136 2.100 1.407 0.942
1.00E-04 3.144 2.105 1.410 0.944
2.20E-04 3.163 2.117 1.417 0.949
4.84E-04 3.204 2.143 1.433 0.958
1.00E-03 3.286 2.193 1.464 0.978
2.20E-03 3.472 2.309 1.537 1.023
4.84E-03 3.868 2.560 1.693 1.120
1.00E-02 4.593 3.029 1.991 1.308
2.20E-02 6.070 4.029 2.648 1.729
4.84E-02 8.579 5.876 3.934 2.589
1.00E-01 11.798 8.541 5.972 4.049
2.20E-01 15.716 12.279 9.226 6.651
4.84E-01 19.213 16.132 13.118 10.241
1.00E+00 21.544 19.069 16.513 13.847

268
To illustrate the impact of utilizing the recommended curves on site

response analyses, a 100-m thick silty sand (SM) deposit was modeled in twenty

six layers and analyzed using the shareware version of ProShake (EduPro, 1998).

A confining-pressure-dependent shear wave velocity, Vs, profile was used (as


shown in Figure 10.10) along with 1500-m/sec Vs at the half space. The Topanga
motion (Maximum Horizontal Acceleration, MHA, = 0.33 g) recorded during the

1994 Northridge earthquake was used as the input “rock” motion.

20

40

Depth, m

60

80

100
0 200 400 600 800 1000
Vs , m/sec

Figure 10.10 Shear wave velocity profile assumed for the 100-m thick silty sand
deposit

269
In Figure 10.11, the acceleration response spectra from two analyses are

presented: 1) using the average generic curves (Seed et al., 1986) to model all

layers, and 2) using the recommended nonlinear curves interpolated for each soil

layer. The response spectrum of the input motion is also shown in this figure. The

response spectra indicate that the recommended nonlinear curves produce an

MHA much higher than that predicted by the average generic curves (0.54 g vs.

0.37 g). Additionally, larger spectral accelerations (typically 30 % to 50 % higher)

are calculated at all periods less than 1 sec for the analysis utilizing the

recommended nonlinear curves.


As discussed in Darendeli et al. (2001) the impact of utilizing a family of

confining-pressure-dependent curves is expected to be more pronounced for

deeper sites subjected to higher intensity input motions due to lower damping

introduced by the confining-pressure-dependent curves. At longer spectral periods

(T > 1 sec), the response is dominated by the overall stiffness of the site. As a

result, the confining-pressure-dependent analyses may tend to predict a smaller

response at longer periods due to the more linear response modeled by these
curves.

270
2.5
This Study (a family of mean curves for PI = 0 %)
Seed et al., 1986 (mean curve for sands)
Input Motion

2.0
Spectral Acceleration, Sa , g

1.5

1.0

0.5

5 % Damping

0.0
0.01 0.1 1 10
Period, T, sec

Figure 10.11 An example of utilizing the recommended normalized modulus


reduction and material damping curves and its impact on estimated
nonlinear site response

271
10.5 SUMMARY

In this chapter, recommended normalized modulus reduction and material

damping curves are presented for soils with a broad range of plasticities confined

over a broad range of mean effective stresses.

The impact of utilizing the recommended curves when assigning nonlinear


soil properties in site response analyses is illustrated by analyzing a 100-m thick

silty sand (SM) deposit using average generic curves (Seed et al., 1986) to model
all twenty six layers, and the recommended nonlinear curves interpolated for each

soil layer. Larger spectral accelerations (typically 30 % to 50 % higher) are

calculated at all periods less than 1 sec for the analysis utilizing the recommended

nonlinear curves than those calculated for the analysis utilizing average generic

curves.

272
CHAPTER 11
UNCERTAINTY ASSOCIATED WITH THE MODEL
PREDICTIONS

11.1 INTRODUCTION

In this chapter, uncertainty associated with the normalized modulus

reduction and material damping curves predicted by the calibrated model is

briefly discussed.
Calculation of standard deviation associated with a point estimate of

normalized shear modulus or material damping ratio, and the covariance structure

of the predicted curves are presented.


Utilization of the calibrated model in probabilistic seismic hazard

assessment is also discussed herein. Integration of random shear-wave velocity

profiles and normalized modulus reduction and material damping curves into

ground motion analysis is recommended. An example regarding incorporation of

the modeled uncertainty (in nonlinear soil behavior) into site response analysis is

also presented.

11.2 UNCERTAINTY IN NONLINEAR SOIL BEHAVIOR

As presented in Chapter Nine, the calibrated model can be utilized to

construct normalized modulus reduction and material damping curves for various

soil types and loading conditions by using the updated mean values of model

parameters (φ1 through φ12) presented in Table 8.12. However, these predicted

273
curves represent average nonlinear curves and the actual data fall into a band of

scatter around these estimates.

At this point, it is important to note that there are two sources of

uncertainty associated with the predicted curves. First, there is uncertainty in the

values of the model parameters. As shown in Table 8.12, where a comparison of

the prior and updated variances of the model parameters are presented, this

component reduces significantly upon calibration of the model.

The second source of uncertainty is the modeled variability (discussed in

Section 7.4.2) of the physical phenomenon. In order to analyze this component of


uncertainty, some of the model parameters (φ13 through φ18) were utilized in
defining the standard deviation and covariance structure of the data (Section
7.4.2) and were simultaneously calibrated utilizing the First-Order, Second-

Moment Bayesian Method.

Table 11.1 presents the predicted mean values and standard deviations

while Table 11.2 presents the covariance structure for a nonplastic soil confined at

1 atm mean effective stress and loaded with ten cycles at 1 Hz accounting for both
components of uncertainty. Tables 11.3 and 11.4 show the same information

predicted by only accounting for modeled variability.

274
Table 11.1 Predicted mean values and standard deviations accounting for
uncertainty in the values of model parameters and variability due to
modeled uncertainty

Point Normalized Shear Modulus Material Damping Ratio, %


Shearing Strain, %
Number Mean Standard Deviation Mean Standard Deviation
1 1.00E-05 0.99945 0.01836 0.80434 0.70766
2 2.20E-05 0.99896 0.01979 0.80816 0.70930
3 4.84E-05 0.99759 0.02254 0.81958 0.71422
4 1.00E-04 0.99546 0.02553 0.83859 0.72232
5 2.20E-04 0.99067 0.03026 0.88404 0.74132
6 4.84E-04 0.98108 0.03683 0.98172 0.78058
7 1.00E-03 0.96353 0.04523 1.17383 0.85250
8 2.20E-03 0.92753 0.05699 1.60166 0.99402
9 4.84E-03 0.86113 0.07116 2.47409 1.23299
10 1.00E-02 0.76094 0.08438 3.95297 1.55606
11 2.20E-02 0.60663 0.09454 6.57873 2.00485
12 4.84E-02 0.42765 0.09556 10.18430 2.49242
13 1.00E-01 0.27720 0.08784 13.78800 2.89875
14 2.20E-01 0.15670 0.07407 17.19890 3.23658
15 4.84E-01 0.08259 0.05961 19.56450 3.45147
16 1.00E+00 0.04417 0.04818 20.71610 3.55137

275
Table 11.2 Predicted covariance structure accounting for uncertainty in the
values of model parameters and variability due to modeled
uncertainty

Point
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Number
1 1.00 0.95 0.90 0.86 0.82 0.78 0.75 0.71 0.68 0.65 0.62 0.59 0.56 0.54 0.51 0.49
2 0.95 1.00 0.95 0.91 0.86 0.82 0.79 0.75 0.71 0.68 0.65 0.62 0.59 0.56 0.54 0.51
3 0.90 0.95 1.00 0.95 0.91 0.86 0.83 0.79 0.75 0.72 0.68 0.65 0.62 0.59 0.56 0.54
4 0.86 0.91 0.95 1.00 0.95 0.90 0.86 0.82 0.78 0.75 0.71 0.68 0.65 0.62 0.59 0.56
5 0.82 0.86 0.91 0.95 1.00 0.95 0.91 0.86 0.82 0.79 0.75 0.71 0.68 0.65 0.62 0.59
6 0.78 0.82 0.86 0.90 0.95 1.00 0.95 0.91 0.86 0.83 0.79 0.75 0.72 0.68 0.65 0.62
7 0.75 0.79 0.83 0.86 0.91 0.95 1.00 0.95 0.90 0.86 0.82 0.78 0.75 0.71 0.68 0.65
8 0.71 0.75 0.79 0.82 0.86 0.91 0.95 1.00 0.95 0.91 0.86 0.82 0.79 0.75 0.71 0.68
9 0.68 0.71 0.75 0.78 0.82 0.86 0.90 0.95 1.00 0.95 0.91 0.86 0.83 0.79 0.75 0.72
10 0.65 0.68 0.72 0.75 0.79 0.83 0.86 0.91 0.95 1.00 0.95 0.90 0.86 0.82 0.78 0.75
11 0.62 0.65 0.68 0.71 0.75 0.79 0.82 0.86 0.91 0.95 1.00 0.95 0.91 0.86 0.82 0.79
12 0.59 0.62 0.65 0.68 0.71 0.75 0.78 0.82 0.86 0.90 0.95 1.00 0.95 0.91 0.86 0.83
13 0.56 0.59 0.62 0.65 0.68 0.72 0.75 0.79 0.83 0.86 0.91 0.95 1.00 0.95 0.90 0.86
14 0.54 0.56 0.59 0.62 0.65 0.68 0.71 0.75 0.79 0.82 0.86 0.91 0.95 1.00 0.95 0.91
15 0.51 0.54 0.56 0.59 0.62 0.65 0.68 0.71 0.75 0.78 0.82 0.86 0.90 0.95 1.00 0.95
16 0.49 0.51 0.54 0.56 0.59 0.62 0.65 0.68 0.72 0.75 0.79 0.83 0.86 0.91 0.95 1.00

276
Table 11.3 Predicted mean values and standard deviations accounting only for
variability due to modeled uncertainty

Point Normalized Shear Modulus Material Damping Ratio, %


Shearing Strain, %
Number Mean Standard Deviation Mean Standard Deviation
1 1.00E-05 0.99945 0.01836 0.80434 0.70507
2 2.20E-05 0.99887 0.02003 0.80892 0.70705
3 4.84E-05 0.99766 0.02243 0.81897 0.71139
4 1.00E-04 0.99546 0.02553 0.83859 0.71978
5 2.20E-04 0.99067 0.03026 0.88404 0.73885
6 4.84E-04 0.98094 0.03692 0.98321 0.77882
7 1.00E-03 0.96353 0.04523 1.17383 0.85036
8 2.20E-03 0.92753 0.05699 1.60166 0.99218
9 4.84E-03 0.86113 0.07116 2.47409 1.23151
10 1.00E-02 0.76093 0.08438 3.95297 1.55488
11 2.20E-02 0.60663 0.09454 6.57873 2.00394
12 4.84E-02 0.42765 0.09556 10.18426 2.49168
13 1.00E-01 0.27720 0.08784 13.78804 2.89811
14 2.20E-01 0.15670 0.07406 17.19895 3.23601
15 4.84E-01 0.08259 0.05961 19.56452 3.45093
16 1.00E+00 0.04417 0.04818 20.71612 3.55085

277
Table 11.4 Predicted covariance structure accounting only for variability due to
modeled uncertainty

Point
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Number
1 1.00 0.95 0.90 0.86 0.82 0.78 0.75 0.71 0.68 0.65 0.62 0.59 0.56 0.54 0.51 0.49
2 0.95 1.00 0.95 0.91 0.86 0.82 0.79 0.75 0.71 0.68 0.65 0.62 0.59 0.56 0.54 0.51
3 0.90 0.95 1.00 0.95 0.91 0.86 0.83 0.79 0.75 0.72 0.68 0.65 0.62 0.59 0.56 0.54
4 0.86 0.91 0.95 1.00 0.95 0.90 0.86 0.82 0.78 0.75 0.71 0.68 0.65 0.62 0.59 0.56
5 0.82 0.86 0.91 0.95 1.00 0.95 0.91 0.86 0.82 0.79 0.75 0.71 0.68 0.65 0.62 0.59
6 0.78 0.82 0.86 0.90 0.95 1.00 0.95 0.91 0.86 0.83 0.79 0.75 0.72 0.68 0.65 0.62
7 0.75 0.79 0.83 0.86 0.91 0.95 1.00 0.95 0.90 0.86 0.82 0.78 0.75 0.71 0.68 0.65
8 0.71 0.75 0.79 0.82 0.86 0.91 0.95 1.00 0.95 0.91 0.86 0.82 0.79 0.75 0.71 0.68
9 0.68 0.71 0.75 0.78 0.82 0.86 0.90 0.95 1.00 0.95 0.91 0.86 0.83 0.79 0.75 0.72
10 0.65 0.68 0.72 0.75 0.79 0.83 0.86 0.91 0.95 1.00 0.95 0.90 0.86 0.82 0.78 0.75
11 0.62 0.65 0.68 0.71 0.75 0.79 0.82 0.86 0.91 0.95 1.00 0.95 0.91 0.86 0.82 0.79
12 0.59 0.62 0.65 0.68 0.71 0.75 0.78 0.82 0.86 0.90 0.95 1.00 0.95 0.91 0.86 0.83
13 0.56 0.59 0.62 0.65 0.68 0.72 0.75 0.79 0.83 0.86 0.91 0.95 1.00 0.95 0.90 0.86
14 0.54 0.56 0.59 0.62 0.65 0.68 0.71 0.75 0.79 0.82 0.86 0.91 0.95 1.00 0.95 0.91
15 0.51 0.54 0.56 0.59 0.62 0.65 0.68 0.71 0.75 0.78 0.82 0.86 0.90 0.95 1.00 0.95
16 0.49 0.51 0.54 0.56 0.59 0.62 0.65 0.68 0.72 0.75 0.79 0.83 0.86 0.91 0.95 1.00

278
The second set of tables (Tables 11.3 and 11.4) can be obtained by

replacing the model parameters in the equations presented in Section 7.4 with

updated mean values in Table 8.12. However, the first set of tables (Tables 11.1

and 11.2) requires a relatively complicated procedure that incorporates the

updated variance of the model parameters in Table 8.12, updated covariance

structure of the model parameters presented in Table 8.13, and derivatives of

equations utilized in modeling mean values and covariance structure presented in

Section 7.4 with respect to each model parameter. The details of this procedure
are beyond the scope of this study and can be found in Ang and Tang (1990).

The comparison of Tables 11.1 and 11.3 indicates that uncertainty in the

value of model parameters has a negligible impact on point estimates. The errors

introduced by calculating mean values and standard deviations without

accounting for uncertainty in the model parameters are less than about 0.1 % for

the mean values and less than about 1 % for the standard deviations. As a result,

the equations presented in Section 7.4 can be used in calculation of mean values
and standard deviations without introducing significant error due to ignoring

uncertainty regarding the model parameters.


Figure 11.1 shows the predicted mean normalized modulus reduction and

material damping curves and standard deviations of the point estimates tabulated

in Table 11.1. These mean curves are identical to the recommended curves

presented in Chapter Ten for deterministic design applications.

279
1.2
(a)
1.0
0.8 G/Gmax Prediction
0.6 σo' = 1 atm,
G/Gmax PI = 0 %,
0.4 N = 10 cycles,
0.2 f = 1 Hz,
OCR = 1
0.0

25
Material Damping Prediction
20 σo' = 1 atm,
PI = 0 %,
15 N = 10 cycles,
D, % f = 1 Hz,
10 OCR = 1

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 11.1 Mean values and standard deviations associated with the point
estimates of (a) normalized modulus reduction and (b) material
damping curves

280
Besides calculating mean normalized modulus reduction and material

damping curves, the additional information generated by analysis of the data

using the Bayesian approach (rather than an ordinary multivariate nonlinear

optimization procedure) can be utilized in predicting nonlinear curves (random

realizations) consistent with the database for probabilistic site response analysis.

As in the case of mean values and standard deviations, the covariance

structures presented in Tables 11.2 and 11.4 are also quite similar. The errors

introduced by calculating covariance structure without accounting for uncertainty

in the model parameters are less than about 0.1 %. As a result, Equation 7.31
presented in Section 7.4.2 can be used in calculation of correlation coefficients

without introducing significant error due to ignoring uncertainty regarding the

model parameters. At the same time, the covariance structure presented in Table

11.2 is unique to the calibrated model and only changes with the number and

relative amplitudes of shearing strains at which normalized modulus reduction

and material damping curves are generated. In other words, as long as nonlinear

curves are generated at the same sixteen shearing strain amplitudes presented in
Table 11.1, the same covariance structure can be utilized regardless of soil type

and loading conditions.


It is also important to note that the covariance structure in Table 11.2 is

not sensitive to the kind of modeled dynamic soil property. In other words, the

covariance structure does not change significantly whether normalized shear

modulus or material damping ratio is being investigated. The difference between

the covariance structure for the normalized modulus reduction curve and the

281
covariance structure for the material damping curve is less than about 0.1 % and

is a result of added uncertainty regarding the additional parameters utilized in

modeling material damping curve.

At the same time, the model calibration was performed assuming no

correlation between normalized modulus reduction and material damping curves,

although the mean G/Gmax and D curves are coupled to each other. In other words,
the calibrated model relates the material damping curve to the normalized

modulus reduction curve in terms of the average estimates, however, where the
point estimates of material damping ratio are relative to the mean material

damping curve is modeled to be independent from where the point estimates of

normalized shear modulus are relative to the mean normalized modulus reduction
curve.

Briefly, correlated random realizations of normalized modulus reduction

and material damping curves can be separately generated utilizing the covariance

structure calculated based on Equation 7.31 and the mean values of the φ17 and φ18
parameters in Table 8.12. The procedure to generate correlated random
realizations consistent with a given model is beyond the scope of this study and

can be found in Ang and Tang (1990).

A realization of normalized modulus reduction and material damping

curves is presented in Figure 11.2 for the same soil type and loading conditions in

Figure 11.1. The mean curves and one standard deviation ranges of normalized
modulus reduction and material damping curves are shown for comparison

purposes. It is important to note that the scatter of the point estimates is not

282
completely random. Any given point estimate is affected by the location of the

neighboring points relative to the mean curve. This is the result of utilizing a

covariance structure in the random realization process.

1.2
(a)
1.0
0.8
0.6
G/Gmax
0.4 Mean Prediction
+/- One Standard Deviation
0.2 Random Realization
0.0

25
Mean Prediction
20 +/- One Standard Deviation
Random Realization
15
D, %
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 11.2 Comparison of the correlated random realization of (a) normalized


modulus reduction and (b) material damping curves relative to the
mean curves and one standard deviation ranges shown in Figure 11.1

283
11.3 UNCERTAINTY IN PREDICTED GROUND MOTIONS DUE TO THE
UNCERTAINTY IN NONLINEAR SOIL BEHAVIOR

Seismic hazard analysis involves the quantitative estimation of ground

shaking at a particular site. If the analysis is carried out assuming a particular

scenario, it is called deterministic seismic hazard analysis. Traditionally, if the

uncertainties in earthquake size, location and time of occurrence are considered in


the analysis, it is called probabilistic seismic hazard analysis. It is important to

note that although probabilistic seismic hazard analysis is site specific,


uncertainties regarding the shear-wave velocity profile have been overlooked in

most cases and uncertainties regarding nonlinear soil behavior at different layers

have been ignored due to lack of data.

It has been established that although the soil profile constitutes a minute

fraction of the travel path from the point of rupture to the ground surface, the
characteristics of the soil layers have a major impact on the amplitude and

frequency content of the ground motion at a geotechnical site. Therefore,

incorporation of uncertainties in soil characteristics into probabilistic seismic


hazard assessment should be expected to result in significant improvement of the

estimated design ground motions.

This study provides the key data, in terms of mean design curves and

uncertainties associated with these curves, required for such an investigation.

However, existing computer programs have to be improved to automatically


incorporate uncertainties in nonlinear soil behavior and shear-wave velocity

profiles in site response analysis. Once such a program becomes available, a

284
number of realizations (varying shear-wave velocity profile and nonlinear curves

at each layer) at a given site utilizing a suite of input motions can be

accomplished with the readily available computational power.

In order to emphasize the impact of modeling uncertainties regarding

nonlinear soil behavior, the example presented in Section 10.4 is reevaluated

using the same input motion and identical shear-wave velocity profile for a

number of realizations.

Due to lack of data, the possible correlation of nonlinear soil behavior

between layers has not been resolved in this study. As a result, two extreme
scenarios regarding the correlation between the twenty five layers within the silty

sand deposit are evaluated: 1) perfectly correlated nonlinear curves, and 2)

completely uncorrelated nonlinear curves. The resulting spectral accelerations

computed using these two extreme cases do not necessarily bound the amplitude

of possible spectral accelerations for design purposes. However, these results

should assist the reader in visualizing the consequences of utilizing different

correlation structures between soil layers.


Figure 11.3 shows the comparison of three spectral accelerations

calculated using perfectly correlated soil layers. The first case is essentially the
same result as shown in Figure 10.11, which was calculated using the mean (µ)
normalized modulus reduction and material damping curves. The other two cases

presented in Figure 11.3 are analyses utilizing normalized modulus reduction and

material damping curves one standard deviation above the mean curves (µ+σ) and
one standard deviation below the mean curves (µ−σ).

285
3.0
Utilizing Mean Curves
Utilizing 1 Standard Deviation Above Average Curves
Utilizing 1 Standard Deviation Below Average Curves

2.5

2.0
Spectral Acceleration, Sa , g

1.5

1.0

0.5

5 % Damping

0.0
0.01 0.1 1 10
Period, T, sec

Figure 11.3 Comparison of spectral accelerations calculated using perfectly


correlated soil layers with µ, µ+σ and µ−σ normalized modulus
reduction and material damping curves

286
Increasing (or decreasing) the stiffness of all layers in the profile (as a

result of simultaneously shifting normalized modulus reduction curves) should be

expected to have a major impact on site period. In the µ+σ case, an increase in
damping accompanied with an increase in stiffness results in a decrease in

estimated ground motion. In the µ−σ case, the modeled normalized modulus
reduction curves are relatively nonlinear and higher strains are generated at deep

layers. Since material damping increases with strain amplitude, more energy is
dissipated at depth, and estimated ground motion turns out to be generally lower

than that estimated using the mean nonlinear curves.

Figure 11.4 also shows comparison of spectral accelerations calculated


using perfectly correlated soil layers with the result calculated using the mean (µ)
curves shown in Figure 10.11. One of the remaining two spectral accelerations in
this figure is computed utilizing µ+σ normalized modulus reduction curve with
µ−σ material damping curve. The third spectral acceleration is computed
utilizing µ−σ normalized modulus reduction curve with µ+σ material damping
curve. This way, if a layer is modeled to be linear relative to the mean curve in
terms of shear modulus, it is also modeled linear relative to the mean curve in

terms of material damping. In the case of the analysis performed on relatively

linear soil layers, changes in the site period can be identified and the spectral

acceleration is observed to be higher at some lower frequencies. Since the other

case involves relatively nonlinear modulus and damping curves, the resulting
ground motions are lower due to increase in energy dissipation.

287
3.0
Utilizing Mean Curves
Utilizing 1 Standard Deviation Linear Curves
Utilizing 1 Standard Deviation Non-Linear Curves

2.5

2.0
Spectral Acceleration, Sa , g

1.5

1.0

0.5

5 % Damping

0.0
0.01 0.1 1 10
Period, T, sec

Figure 11.4 Comparison of spectral accelerations calculated using perfectly


correlated soil layers with 1) µ curves, 2) +σ normalized modulus
reduction and −σ material damping curves, and 3) −σ normalized
modulus reduction and +σ material damping curves

288
Fifty realizations utilizing completely uncorrelated nonlinear curves are

presented in Figure 11.5. Although completely uncorrelated soil layers are

unlikely to be a common scenario, the reader must keep in mind the possibility of

missing a thin soft layer during real-life site investigations.

Since the equivalent linear analysis program (EduPro, 1998) utilized in the

analysis is not designed for random realizations, an internal file that contains the

nonlinear curves had to be modified before each run. Although computation time

required for each run was merely about 10 seconds, modification of this file

significantly slowed the process. A number of improvements in site response


analysis programs will be required if they are to be utilized in probabilistic

seismic assessment as recommended.

In Figure 11.5, significantly different spectral accelerations are presented

for a given input motion and shear-wave velocity profile. A single relatively

nonlinear soil layer was observed to reduce surface motions drastically in some of

these realizations. As a result, graphing all spectral accelerations for a large

number of realizations or analyzing histograms of spectral accelerations at certain


periods (Figures 11.6 and 11.7) is recommended rather than analyzing µ, µ+σ,
or µ+2σ spectral accelerations as shown in Figure 11.8. However, a comparison
of the information presented in Figure 11.8 with the result computed utilizing the
mean curves (Figure 11.9) shows that depending on the consequences of failure,

the design acceleration response spectrum may have to be selected much higher
than the deterministic spectrum (estimated utilizing the mean curves) even though

the input motion and the shear-wave velocity profile were fixed in this example.

289
3.0
Random Realization 5 % Damping

2.5

2.0
Spectral Acceleration, Sa , g

1.5

1.0

0.5

0.0
0.01 0.1 1 10
Period, T, sec

Figure 11.5 Fifty realizations of spectral acceleration computed using completely


uncorrelated soil layers with randomly generated normalized
modulus reduction and material damping curves

290
35
(a)
30
25
Frequency

20
15
10
5
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Acceleration at 0.1 sec Period, g

35
(b)
30
25
Frequency

20
15
10
5
0
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2 2.25 2.5
Acceleration at 0.3 sec Period, g

Figure 11.6 Histograms of spectral accelerations from fifty realizations presented


in Figure 11.5 (a) at 0.1 sec and (b) at 0.3 sec

291
35
(a)
30
25
Frequency

20
15
10
5
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Acceleration at 1 sec Period, g

35
(b)
30
25
Frequency

20
15
10
5
0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
Acceleration at 3 sec Period, g

Figure 11.7 Histograms of spectral accelerations from fifty realizations presented


in Figure 11.5 (a) at 1 sec and (b) at 3 sec

292
3.0
Mean of Random Realizations
Mean +1 Standard Deviation of Random Realizations
Mean +2 Standard Deviation of Random Realizations

2.5

2.0
Spectral Acceleration, Sa , g

1.5

1.0

0.5

5 % Damping
0.0
0.01 0.1 1 10
Period, T, sec

Figure 11.8 Distribution of fifty realizations of spectral acceleration presented in


Figure 11.5

293
3.0
Mean of Random Realizations
Mean +1 Standard Deviation of Random Realizations
Mean +2 Standard Deviation of Random Realizations
Utilizing Mean Curves
2.5

2.0
Spectral Acceleration, Sa , g

1.5

1.0

0.5

5 % Damping
0.0
0.01 0.1 1 10
Period, T, sec

Figure 11.9 Comparison of the spectral accelerations from the fifty realizations
with the results computed utilizing mean normalized modulus
reduction and material damping curves

294
It is important to point out that the discussion and accompanying figures

presented above are related to observations on a single input motion and a

uniform silty sand deposit with increasing shear wave velocity with depth. The

consequences of ignoring the uncertainty in nonlinear soil behavior are expected

to be more pronounced in the case of higher intensity input motions and more

complicated soil deposits.

11.4 SUMMARY

The uncertainty associated with the recommended normalized modulus

reduction and material damping curves is discussed in this chapter. The impact of
such uncertainty on estimated ground motions for a given input motion and shear-

wave velocity profile is presented.

For probabilistic seismic hazard analysis applications, the recommended

procedure for handling the uncertainty in the shear-wave velocity profile and the

nonlinear soil behavior are also discussed.


Since this study provides the only available data regarding uncertainty

associated with the recommended normalized modulus reduction and material

damping curves, the results presented in this chapter are believed to be a very
important contribution to state of the art in geotechnical earthquake engineering.

295
CHAPTER 12
SUMMARY AND CONCLUSIONS

12.1 SUMMARY

In this study, the effects of soil type and loading conditions on dynamic

soil properties (presented in terms of normalized shear modulus and material

damping curves) have been quantified based on the data that has been collected at
the University of Texas at Austin over the past decade. Information regarding the

laboratory testing equipment used to collect the data and a general description of

the properties of the specimens included in the database are presented in Chapters
Two and Three, respectively.

The general trends regarding nonlinear soil behavior observed during the

course of this study and reported in the literature are presented in Chapter Four.

Parameters that control nonlinear soil behavior and their relative importance in

terms of affecting normalized modulus reduction and material damping curves


based on general trends observed during the course of this study are presented in

Table 12.1.
Based on the general trends, the successes and shortcomings of various

empirical relationships utilized in the state of practice are evaluated in Chapter

Five and a four-parameter (reference strain, curvature coefficient, small-strain

material damping ratio and the scaling coefficient) soil model that is capable of

capturing these general trends is proposed in Chapter Six (Equation 12.1).

296
Table 12.1 Parameters that control nonlinear soil behavior and their relative
importance in terms of affecting normalized modulus reduction and
material damping curves based on general trends observed during
the course of this study

Impact on Normalized Impact on Material


Parameter
Modulus Reduction Curve Damping Curve

Strain Amplitude *** ***


Mean Effective Confining
Pressure
*** ***
Soil Type and Plasticity *** ***
Number of Loading Cycles *+ ***++
Frequency of Loading
(above 1 Hz)
* **
Overconsolidation Ratio * *
Void Ratio * *
Degree of Saturation * *
Grain Characteristics, Size,
Shape, Gradation, Mineralogy
* *
+
*** Very Important On competent soils included in this study
++
** Important Soil Type Dependent
* Less Important

G 1
= a
(12.1a)
Gmax γ 
1 +  
γ r 
0.1
 G 
D Adjusted = b *   * DMa sin g + Dmin (12.1b)
 Gmax 

The First-Order, Second-Moment Bayesian Method utilized in the

statistical analysis of the data is discussed in Chapter Seven. An eighteen

parameter model that relates reference strain, curvature coefficient, small-strain

material damping ratio and scaling coefficient to soil type and loading conditions,

297
and that characterizes the covariance structure of the predicted normalized

modulus reduction and material damping curves is also presented in this chapter

(Equation 12.2).
γ r = (φ1 + φ 2 * PI * OCR φ ) * σ o 'φ
3 4
(12.2a)
a = φ5 (12.2b)
Dmin = (φ 6 + φ 7 * PI * OCR φ8 ) * σ o 'φ9 *[1 + φ10 * ln( frq )] (12.2c)

b = φ11 + φ12 * ln( N ) (12.2d)


0.25 (G / Gmax − 0.5) 2
σ NG = exp(φ13 ) + − (12.2e)
exp(φ14 ) exp(φ14 )

σ D = exp(φ15 ) + exp(φ16 ) * D (12.2f)


−1 − ln γ i − ln γ j
ρ i , j = exp( ) * exp( ) (12.2g)
exp(φ17 ) exp(φ18 )

Statistical analysis of various subsets of the data and the model calibration
process are briefly described in Chapter Eight.

Chapters Nine and Ten present the equations, graphical solutions, plots

and tables of recommended normalized modulus reduction and material damping

curves for deterministic site-specific analysis. The proposed curves are also

compared with empirical curves widely accepted in state of practice in Chapter

Nine (as shown in Figures 12.1 and 12.2).

Finally, uncertainty associated with the recommended curves is discussed


in Chapter Eleven.

298
1.2
(a)
1.0
0.8
0.6 G/Gmax Prediction
G/Gmax ( PI = 0 %, N = 10 cycles,
0.4 f = 1 Hz, OCR = 1 )
0.2 Seed et al. (1986)
0.0

25
Material Damping Prediction
20 σo' = 0.25 atm
σo' = 1 atm
15 σo' = 4 atm
D, %
σo' = 16 atm
10

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 12.1 Comparison of the effect of confining pressure on nonlinear soil


behavior of sand (PI = 0 %) predicted by the calibrated model and
empirical curves proposed for sands by Seed et al. (1986)

299
1.2
(a)
1.0
0.8
G/Gmax Prediction
0.6
G/Gmax ( σo' = 1 atm, N = 10 cycles,
0.4 f = 1 Hz, OCR = 1 )

0.2 Vucetic and Dobry (1991)


0.0

25
Material Damping Prediction
20 PI = 0 %
PI = 15 %
15 PI = 30 %
D, % PI = 50 %
10 PI = 100 %

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 12.2 Comparison of the effect of soil plasticity on nonlinear soil behavior
predicted by the calibrated model and empirical curves proposed by
Vucetic and Dobry (1991)

300
12.2 CONCLUSIONS

A new family of normalized modulus reduction and material damping

design curves is proposed utilizing a four-parameter model calibrated to a rather

large database of resonant column and torsional shear test results. One of the

unique features of this study is the consideration for uncertainty associated with
the recommended curves. Figure 12.3 shows mean values predicted using the

calibrated model and uncertainty associated with these point estimates.


This study is believed to be a valuable contribution to the state of the art

because it provides the means to incorporate the uncertainty in nonlinear soil

behavior into probabilistic seismic hazard analysis. However, utilization of this

work requires improvement of site response analysis programs available at this

point in time.

301
1.2
(a)
1.0
0.8 G/Gmax Prediction
0.6 σo' = 1 atm,
G/Gmax PI = 0 %,
0.4 N = 10 cycles,
0.2 f = 1 Hz,
OCR = 1
0.0

25
Material Damping Prediction
20 σo' = 1 atm,
PI = 0 %,
15 N = 10 cycles,
D, % f = 1 Hz,
10 OCR = 1

5
(b)
0
0.0001 0.001 0.01 0.1 1
Shearing Strain, γ ,%

Figure 12.3 Mean values and standard deviations associated with the point
estimates of (a) normalized modulus reduction and (b) material
damping curves

302
APPENDIX A

HEADER FILE

FOR

FIRST ORDER SECOND MOMENT

BAYESIAN ANALYSIS

OF

RESONANT COLUMN

AND

TORSIONAL SHEAR

TEST RESULTS

303
// Modelh.h : Header file for RCTS data
//
class ModelStructure : public NormLike
{
public:
int nCOV;
// Data Indices
int
nlocation,nsoil,nspecimen,nswv,ndisturbance,npressure,ntest,nPI,nOCR,ne,nconp
re,nfrq,
nN,nstr,ncorrstr,nTYPE;
// Model Indices
int iphi1;
int iphi2;
int iphi3;
int iphi4;
int iphi5;
int iphi6;
int iphi7;
int iphi8;
int iphi9;
int iphi10;
int iphi11;
int iphi12;
int istdGa,istdGb,istdDa,istdDb,ithetanugget;
int ntheta;
iarray itheta;
double scalar;

void Initialize(DataStructure &Data, int nPhi, int nconst, double *xconst);


void Feasible(double *x);
void OutputModel(DataStructure &Data, double *x, CString outfilename);
void PrintYMean(DataStructure &Data, double *x, CString output);
void CalculateYMeanC(DataStructure &Data, double *x, darray
&YMeanC, iarray &index);
void CalculatedYMeanCiMM(int iv, DataStructure &Data, double *x,
darray &dYMeani, iarray
&index);
void Calculated2YMeanCijMM(int iv, int jv, DataStructure &Data, double
*x,
darray &dYMeani, darray &dYMeanj, darray &d2YMeanij,
iarray &index);

304
void CalculateYCOVC(DataStructure &Data, double *x, darray &YMean,
CovMatrix &YCOV,
iarray &index);
void CalculatedYCOVCiMM(int iv, DataStructure &Data, double *x,
darray &YMean,
darray &dYMeani,
CovMatrix &YCOV, smatrixsolve &dYCOVi, iarray &index);
void Calculated2YCOVCijMM(int iv, int jv, DataStructure &Data, double
*x, darray &YMean,
darray &dYMeani, darray
&dYMeanj, darray &d2YMeanij, CovMatrix &YCOV,
smatrixsolve &dYCOVi,
smatrixsolve &dYCOVj, smatrixsolve &d2YCOVij, iarray &index);
double CalculateYrhoab(DataStructure &Data, double *x, int ka, int kb);
double YCOVrho(double *tau, double *x);
double dYCOVrhoi(int iv, double *tau, double *x);
double d2YCOVrhoij(int iv, int jv, double *tau, double *x);

};

305
APPENDIX B

MODEL FILE

FOR

FIRST ORDER SECOND MOMENT

BAYESIAN ANALYSIS

OF

RESONANT COLUMN

AND

TORSIONAL SHEAR

TEST RESULTS

306
// Model.cpp : Model for RCTS Data
//
#include "stdafx.h"
#include <afxwin.h>
#include <iostream.h>
#include <fstream.h>
#include <math.h>
#include <direct.h>
#include <time.h>
#include "machh.h"
#include "compareh.h"
#include "dblash.h"
#include "_arrayh.h"
#include "_array2h.h"
#include "_array3h.h"
#include "matrixh.h"
#include "smatrixh.h"
#include "gmatrixh.h"
#include "covmatrixh.h"
#include "dblash.h"
#include "goldenh.h"
#include "rqph.h"
#include "Datah.h"
#include "NormalLikeh.h"
#include "Modelh.h"

void ModelStructure::Initialize(DataStructure &Data, int nPhi, int nconst, double


*xconst)
{
/* Data Structure:
Data.d[0] = location # (1=Northern CA,
2=Southern CA,
3=South Carolina,
4=Lotung, Taiwan)
Data.d[1] = soil type ( 1=sands with fines < 12%,
2=sands with fines > 12% [<
50%],
3=silts,
4=clays)
Data.d[2] = specimen # (3 or 4 digit specimen ID
UTA-1-C=103, UT-24-
F=2406)

307
Data.d[3] = in-situ shear wave velocity (not utilized in this study,

a value of 500 is assigned

to all specimens)
Data.d[4] = shear wave velocity ratio (not utilized in this study,

a value of 1.0 is assigned

to all specimens)
Data.d[5] = pressure # (some specimens are tested
at multiple confining
pressures)
Data.d[6] = test type ( 1.05-1.2=LA_TS,
2=HA_TS1,
3=HA_TS10,
4=HA_RC)
Data.d[7] = plasticity index, PI (%)
Data.d[8] = overconsolidation ratio, OCR
Data.d[9] = void ratio, e
Data.d[10] = isotropic effective confining pressure, conpre (atm)
Data.d[11] = loading frequency, frq (Hz)
Data.d[12] = number of loading cycles, N
Data.d[13] = peak strain for modulus, str (%)
Data.d[14] = corrected strain for damping (diffrent only in high amplitude
RC), corr_str (%)
Data.d[15] = indicator of data type ( 0 for normalized modulus, NG

1 for material damping ratio, D)


Data.d[16] = experimental observation ( normalized modulus, NG or

material damping ratio, D, %)*/

nlocation = 0; /* index in Data.d of sample


location */
nsoil = nlocation + 1; /* index in Data.d of soil type */
nspecimen = nsoil + 1; /* index in Data.d of
specimen # */
nswv = nspecimen + 1; /* index in Data.d of in-situ
shear wave velocity */
ndisturbance = nswv + 1; /* index in Data.d of shear wave
velocity ratio (disturbance) */

308
npressure = ndisturbance + 1; /* index in Data.d of pressure # */
ntest = npressure + 1; /* index in Data.d of test type */
nPI = ntest + 1; /* index in Data.d of PI */
nOCR = nPI + 1; /* index in Data.d of
OCR */
ne = nOCR + 1; /* index in Data.d of e
*/
nconpre = ne + 1; /* index in Data.d of conpre
*/
nfrq = nconpre + 1; /* index in Data.d of frq */
nN = nfrq + 1; /* index in Data.d of N */
nstr = nN + 1; /* index in Data.d of str */
ncorrstr = nstr + 1; /* index in Data.d of corr_str */
nTYPE = ncorrstr + 1; /* index in Data.d of Data
Type (G versus D) */
//-------------------------------------------
censorcheck = 0; /* 0 = all point measurments, 1 = some censored
measurements */
ncflag = 99; /* index in Data.d of censor flag (indicates if measurement is
censored, if censorcheck = 1) */
nydown = 99; /* index in Data.d of lower bound measurements, if
censorcheck = 1 */
ny = nTYPE + 1; /* index in Data.d of point/upper bound measurements
*/
ny0 = 0; /* index of first useable data point */
//-------------------------------------------
nx = nPhi; /* total number of parameters */
nCOV = nx; /* number of separate (not also used for mean) variance
parameters */
nmu = nPhi - nCOV; /* number of separate (not also used for variance)
mean parameters */

/* Indices in ModelStructure for the model parameters */


iphi1 = 0;
iphi2 = iphi1 + 1;
iphi3 = iphi2 + 1;
iphi4 = iphi3 + 1;
iphi5 = iphi4 + 1;
iphi6 = iphi5 + 1;
iphi7 = iphi6 + 1;
iphi8 = iphi7 + 1;
iphi9 = iphi8 + 1;

309
iphi10 = iphi9 + 1;
iphi11 = iphi10 + 1;
iphi12 = iphi11 + 1;

istdGa = iphi12 + 1; /* ln */
istdGb = istdGa + 1; /* ln */
istdDa = istdGb + 1; /* ln */
istdDb = istdDa + 1; /* ln */

ithetanugget = istdDb + 1; /* ln */
ntheta = 5; /* ln */
itheta.construct(ntheta);
itheta[0] = ithetanugget + 1;
itheta[1] = itheta[0] + 1;
itheta[2] = itheta[1] + 1;
itheta[3] = itheta[2] + 1;
itheta[4] = itheta[3] + 1;

/* Scale measurements to avoid numerical precision problems in calculations */


scalar = 1.0;

if (censorcheck == 1) dscal(Data.nMeas,scalar,Data.d[nydown],1);
dscal(Data.nMeas,scalar,Data.d[ny],1);

rhozero = 1.0e-2; /* effective zero value for correlation */


}

void ModelStructure::Feasible(double *x)


{}

void ModelStructure::OutputModel(DataStructure &Data, double *x, CString


outfilename)
{
PrintYMean(Data,x,outfilename);
}

310
APPENDIX C

FILE USED IN ESTIMATING MEAN VALUES

FOR

FIRST ORDER SECOND MOMENT

BAYESIAN ANALYSIS

OF

RESONANT COLUMN

AND

TORSIONAL SHEAR

TEST RESULTS

311
// RCTSYMean.cpp : Proposed Equations for RCTS data
//
#include "stdafx.h"
#include <afxwin.h>
#include <iostream.h>
#include <fstream.h>
#include <math.h>
#include <direct.h>
#include "machh.h"
#include "compareh.h"
#include "dblash.h"
#include "_arrayh.h"
#include "_array2h.h"
#include "_array3h.h"
#include "matrixh.h"
#include "smatrixh.h"
#include "gmatrixh.h"
#include "covmatrixh.h"
#include "dblash.h"
#include "goldenh.h"
#include "rqph.h"
#include "Datah.h"
#include "NormalLikeh.h"
#include "Modelh.h"

void ModelStructure::PrintYMean(DataStructure &Data, double *x, CString


output)
{
int n,k;
int ytype;
double yk;

n = Data.nMeas;
darray YMean(n);

iarray index(n);
for (k = 0; k < n; k++) index[k] = k;

CalculateYMeanC(Data,x,YMean,index);
CovMatrix YCOV(n);
CalculateYCOVC(Data,x,YMean,YCOV,index);
ofstream out(output);

312
out << "Point" << "\t" << "ytype" << "\t" << "yk" << "\t" << "YMean" <<
"\t" << "YCOV" "\n";

for (k = 0; k < n; k++)


{
yk = Data.d[ny][k];
ytype = int(Data.d[nTYPE][k]);
out << k << "\t" << ytype << "\t" << yk/scalar << "\t" <<
YMean[k]/scalar << "\t" << YCOV.G.xptr[k] << "\n";
}

out.close();
}

void ModelStructure::CalculateYMeanC(DataStructure &Data, double *x, darray


&YMeanC, iarray &index)
{
int kindex,nindex,k;
double dbl_par_phi[25];// par stand for model parameters
double dbl_atr_str, dbl_atr_corr_str, dbl_atr_PI, dbl_atr_OCR,
dbl_atr_e, dbl_atr_Fe, dbl_atr_conpre, dbl_atr_frq, dbl_atr_N; // atr stands
for attributes
double dbl_refstr, dbl_a, dbl_NG, dbl_NG_corrstr, dbl_DMasing,
dbl_c1, dbl_c2, dbl_c3, dbl_Dmin, dbl_b, dbl_D; // Dependent
intermediate variables
int datatype;
double dbl_con_pi = 3.1415926535; //Constant PI

nindex = index.n; // total number of nearby data points

// assign values to model parameters


dbl_par_phi[1] = x[iphi1];
dbl_par_phi[2] = x[iphi2];
dbl_par_phi[3] = x[iphi3];
dbl_par_phi[4] = x[iphi4];
dbl_par_phi[5] = x[iphi5];
dbl_par_phi[6] = x[iphi6];
dbl_par_phi[7] = x[iphi7];
dbl_par_phi[8] = x[iphi8];
dbl_par_phi[9] = x[iphi9];
dbl_par_phi[10] = x[iphi10];
dbl_par_phi[11] = x[iphi11];

313
dbl_par_phi[12] = x[iphi12];

// Step through nearby data set and get YMean for each point
for (kindex = 0; kindex < nindex; kindex++)
{
k = index[kindex];
// assign values to attributes
dbl_atr_str = Data.d[nstr][k];
dbl_atr_corr_str = Data.d[ncorrstr][k];
dbl_atr_PI = Data.d[nPI][k];
dbl_atr_OCR = Data.d[nOCR][k];
dbl_atr_e = Data.d[ne][k];
dbl_atr_conpre = Data.d[nconpre][k];
dbl_atr_frq = Data.d[nfrq][k];
dbl_atr_N = Data.d[nN][k];
datatype = int(Data.d[nTYPE][k]);

// calculation of normalized modulus and damping values


// for given atributes and model parameters

// although the effect of void ratio is not accounted for in this study
// the code is written so that an F(e) term such as the one below
// can be included in the future
// dbl_atr_Fe=0.3 + 0.7 * pow (dbl_atr_e,2);

dbl_atr_Fe=1.0;

dbl_refstr=(dbl_par_phi[1]+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OC
R,dbl_par_phi[3]))

*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4]);

dbl_a=dbl_par_phi[5];

dbl_NG=1.0/(1+pow((dbl_atr_str/dbl_refstr),dbl_a));

dbl_NG_corrstr=1.0/(1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a));

dbl_DMasing=(100.0/dbl_con_pi)*(4*(dbl_atr_corr_str-
dbl_refstr*log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr))

314
/(pow(dbl_atr_corr_str,2)/(dbl_atr_corr_str+dbl_refstr))-2);

dbl_c1= -1.1143*pow(dbl_a,2)+1.8618*dbl_a+0.2523;

dbl_c2= 0.0805*pow(dbl_a,2)-0.0710*dbl_a-0.0095;

dbl_c3= -0.0005*pow(dbl_a,2)+0.0002*dbl_a+0.0003;

dbl_Dmin=
(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8]))

*pow(dbl_atr_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*log(dbl_atr_fr
q));

dbl_b=dbl_par_phi[11]+dbl_par_phi[12]*log(dbl_atr_N);

dbl_D=dbl_Dmin+dbl_b*pow(dbl_NG_corrstr,0.1)*(dbl_c1*dbl_DMasin
g+dbl_c2*pow(dbl_DMasing,2)
+dbl_c3*pow(dbl_DMasing,3));

if (datatype == 0)
{
YMeanC[kindex] = scalar*dbl_NG;
}
else
{
YMeanC[kindex] = scalar*dbl_D;
}

} // for (kindex = 0; kindex < nindex; kindex++)

void ModelStructure::CalculatedYMeanCiMM(int iv, DataStructure &Data,


double *x,
darray
&dYMeani, iarray &index)
{
int k,kindex,nindex;

315
double dbl_par_phi[25], dbl_par_dphi[25]; // par stand for model
parameters
double dbl_atr_str, dbl_atr_corr_str, dbl_atr_PI, dbl_atr_OCR,
dbl_atr_e, dbl_atr_Fe, dbl_atr_conpre, dbl_atr_frq, dbl_atr_N; // atr stands
for attributes
double dbl_refstr, dbl_a, dbl_NG, dbl_NG_corrstr, dbl_DMasing,
dbl_c1, dbl_c2, dbl_c3, dbl_Dmin, dbl_b, dbl_D; // Dependent
intermediate variables
double dbl_drefstr, dbl_da, dbl_dNG, dbl_dNG_corrstr, dbl_dc1, dbl_dc2,
dbl_dc3, dbl_dDMasing,
dbl_dDmin, dbl_db, dbl_dD; // First Order Partial Derivative of
dependent intermediate variables
int datatype;
int int_loopcounter_i,int_loopcounter_j;
double dbl_con_pi = 3.1415926535; //Constant PI

nindex = index.n;

// assign values to model parameters


dbl_par_phi[1] = x[iphi1];
dbl_par_phi[2] = x[iphi2];
dbl_par_phi[3] = x[iphi3];
dbl_par_phi[4] = x[iphi4];
dbl_par_phi[5] = x[iphi5];
dbl_par_phi[6] = x[iphi6];
dbl_par_phi[7] = x[iphi7];
dbl_par_phi[8] = x[iphi8];
dbl_par_phi[9] = x[iphi9];
dbl_par_phi[10] = x[iphi10];
dbl_par_phi[11] = x[iphi11];
dbl_par_phi[12] = x[iphi12];

// set all dphi equal to zero


for (int_loopcounter_j=0;int_loopcounter_j<nx;int_loopcounter_j++)
{
dbl_par_dphi[int_loopcounter_j]=0;
}

// calculation of first order derivatives for iv


int_loopcounter_i=iv+1;

// calculate derivatives for phi[int_loopcounter_i]

316
dbl_par_dphi[int_loopcounter_i]=1;

for (kindex = 0; kindex < nindex; kindex++)


{
k = index[kindex];

// assign values to attributes


dbl_atr_str= Data.d[nstr][k];
dbl_atr_corr_str= Data.d[ncorrstr][k];
dbl_atr_PI= Data.d[nPI][k];
dbl_atr_OCR= Data.d[nOCR][k];
dbl_atr_e= Data.d[ne][k];
dbl_atr_conpre= Data.d[nconpre][k];
dbl_atr_frq= Data.d[nfrq][k];
dbl_atr_N= Data.d[nN][k];
datatype = int(Data.d[nTYPE][k]);

// calculation of normalized modulus and damping values


// for given atributes and model parameters

// although the effect of void ratio is not accounted for in this study
// the code is written so that an F(e) term such as the one below
// can be included in the future
// dbl_atr_Fe=0.3 + 0.7 * pow (dbl_atr_e,2);

dbl_atr_Fe=1.0;

dbl_refstr=(dbl_par_phi[1]+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OC
R,dbl_par_phi[3]))

*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4]);

dbl_a=dbl_par_phi[5];

dbl_NG=1.0/(1+pow((dbl_atr_str/dbl_refstr),dbl_a));

dbl_NG_corrstr=1.0/(1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a));

dbl_DMasing=(100.0/dbl_con_pi)*(4*(dbl_atr_corr_str-
dbl_refstr*log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr))

317
/(pow(dbl_atr_corr_str,2)/(dbl_atr_corr_str+dbl_refstr))-2);

dbl_c1= -1.1143*pow(dbl_a,2)+1.8618*dbl_a+0.2523;

dbl_c2= 0.0805*pow(dbl_a,2)-0.0710*dbl_a-0.0095;

dbl_c3= -0.0005*pow(dbl_a,2)+0.0002*dbl_a+0.0003;

dbl_Dmin=
(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8]))

*pow(dbl_atr_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*log(dbl_atr_fr
q));

dbl_b=dbl_par_phi[11]+dbl_par_phi[12]*log(dbl_atr_N);

dbl_D=dbl_Dmin+dbl_b*pow(dbl_NG_corrstr,0.1)*(dbl_c1*dbl_DMasin
g+dbl_c2*pow(dbl_DMasing,2)
+dbl_c3*pow(dbl_DMasing,3));

//"FIRST ORDER DERIVATIVES" CODE FOR phi[int_loopcounter_i]


//-------------------------------------------------------

dbl_drefstr=dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*dbl_par_dp
hi[1]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*dbl_atr_Fe*pow(dbl_atr
_conpre,dbl_par_phi[4])*dbl_par_dphi[2]

+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*log(dbl_
atr_OCR)*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*dbl_par_dphi[3]

+(dbl_par_phi[1]+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_
phi[3]))*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*log(dbl_atr_conpre)*d
bl_par_dphi[4];

dbl_da=dbl_par_dphi[5];

318
dbl_dNG=(1/pow((1+pow((dbl_atr_str/dbl_refstr),dbl_a)),2))*pow((dbl_a
tr_str/dbl_refstr),dbl_a)*(dbl_a/dbl_refstr)*dbl_drefstr
+(-
1/pow((1+pow((dbl_atr_str/dbl_refstr),dbl_a)),2))*pow((dbl_atr_str/dbl_refstr),db
l_a)*log(dbl_atr_str/dbl_refstr)*dbl_da;

dbl_dNG_corrstr=(1/pow((1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a)),2))
*pow((dbl_atr_corr_str/dbl_refstr),dbl_a)*(dbl_a/dbl_refstr)*dbl_drefstr
+(-
1/pow((1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a)),2))*pow((dbl_atr_corr_str/db
l_refstr),dbl_a)*log(dbl_atr_corr_str/dbl_refstr)*dbl_da;

dbl_dc1=(-2.2286*dbl_a+1.8618)*dbl_da;

dbl_dc2=( 0.1610*dbl_a-0.0710)*dbl_da;

dbl_dc3=(-0.0010*dbl_a+0.0002)*dbl_da;

dbl_dDMasing=(-
400)*((log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr)*dbl_atr_corr_str

+2*dbl_refstr*log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr)
-
2*dbl_atr_corr_str)/(pow(dbl_atr_corr_str,2)*dbl_con_pi))
*dbl_drefstr;

dbl_dDmin=pow(dbl_atr_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*log
(dbl_atr_frq))*dbl_par_dphi[6]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*pow(dbl_atr_conpre,dbl_par_ph
i[9])*(1+dbl_par_phi[10]*log(dbl_atr_frq))*dbl_par_dphi[7]

+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*log(dbl_atr_OC
R)*pow(dbl_atr_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*log(dbl_atr_frq))*d
bl_par_dphi[8]

+(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])
)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_conpre)*(1+dbl_par_phi[10]*
log(dbl_atr_frq))*dbl_par_dphi[9]

319
+(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])
)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_frq)*dbl_par_dphi[10];

dbl_db=dbl_par_dphi[11]+log(dbl_atr_N)*dbl_par_dphi[12];

dbl_dD=dbl_dDmin

+pow(dbl_NG_corrstr,0.1)*dbl_DMasing*(dbl_c1+dbl_c2*dbl_DMasing+dbl_c3
*pow(dbl_DMasing,2))*dbl_db

+0.1*dbl_b*dbl_DMasing*((dbl_c1+dbl_c2*dbl_DMasing+dbl_c3*pow(dbl_DM
asing,2))/pow(dbl_NG_corrstr,0.9))*dbl_dNG_corrstr

+dbl_b*pow(dbl_NG_corrstr,0.1)*(dbl_c1+2*dbl_c2*dbl_DMasing+3*dbl_c3*p
ow(dbl_DMasing,2))*dbl_dDMasing

+dbl_b*pow(dbl_NG_corrstr,0.1)*dbl_DMasing*dbl_dc1

+dbl_b*pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,2)*dbl_dc2

+dbl_b*pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,3)*dbl_dc3;
//-------------------------------------------------------

if (datatype == 0)
{
dYMeani[kindex] = scalar*dbl_dNG;
}
else
{
dYMeani[kindex] = scalar*dbl_dD;
}

} // for (kindex = 0; kindex < nindex; kindex++)

void ModelStructure::Calculated2YMeanCijMM(int iv, int jv, DataStructure


&Data, double *x,

320
darray &dYMeani, darray &dYMeanj, darray
&d2YMeanij, iarray &index)
{
int k,kindex,nindex;
double dbl_par_phi[25], dbl_par_diphi[25], dbl_par_djphi[25]; // par stand
for model parameters
double dbl_atr_str, dbl_atr_corr_str, dbl_atr_PI, dbl_atr_OCR,
dbl_atr_e, dbl_atr_Fe, dbl_atr_conpre, dbl_atr_frq, dbl_atr_N; // atr stands
for attributes
double dbl_refstr, dbl_a, dbl_NG, dbl_NG_corrstr, dbl_DMasing,
dbl_c1, dbl_c2, dbl_c3, dbl_Dmin, dbl_b, dbl_D; // Dependent
intermediate variables
double dbl_direfstr, dbl_dia, dbl_diNG, dbl_diNG_corrstr, dbl_dic1,
dbl_dic2, dbl_dic3, dbl_diDMasing,
dbl_diDmin, dbl_dib, dbl_diD; // diNG stands for first order partial
derivative of NG with respect to phi[i]
double dbl_djrefstr, dbl_dja, dbl_djNG, dbl_djNG_corrstr, dbl_djc1,
dbl_djc2, dbl_djc3, dbl_djDMasing,
dbl_djDmin, dbl_djb, dbl_djD; // djNG stands for first order partial
derivative of NG with respect to phi[j]
double dbl_d2refstr, dbl_d2a, dbl_d2NG, dbl_d2NG_corrstr, dbl_d2c1,
dbl_d2c2, dbl_d2c3, dbl_d2DMasing,
dbl_d2Dmin, dbl_d2b, dbl_d2D; // d2NG stands for second order
partial derivative of NG with respect to phi[i]=phi[j]
int datatype;
int
int_loopcounter_k,int_loopcounter_l,int_loopcounter_m,int_loopcounter_n;
double dbl_con_pi = 3.1415926535; //Constant PI

nindex = index.n;

// assign values to model parameters


dbl_par_phi[1] = x[iphi1];
dbl_par_phi[2] = x[iphi2];
dbl_par_phi[3] = x[iphi3];
dbl_par_phi[4] = x[iphi4];
dbl_par_phi[5] = x[iphi5];
dbl_par_phi[6] = x[iphi6];
dbl_par_phi[7] = x[iphi7];
dbl_par_phi[8] = x[iphi8];
dbl_par_phi[9] = x[iphi9];
dbl_par_phi[10] = x[iphi10];

321
dbl_par_phi[11] = x[iphi11];
dbl_par_phi[12] = x[iphi12];

// set all diphi equal to zero


for (int_loopcounter_l=0;int_loopcounter_l<nx;int_loopcounter_l++)
{
dbl_par_diphi[int_loopcounter_l]=0;
}

// set all djphi equal to zero


for (int_loopcounter_n=0;int_loopcounter_n<nx;int_loopcounter_n++)
{
dbl_par_djphi[int_loopcounter_n]=0;
}

// calculation of second order derivatives


(d/diphi[int_loopcounter_k]*djphi[int_loopcounter_m])

// calculation of second order derivatives for iv, jv


int_loopcounter_k=iv+1;
int_loopcounter_m=jv+1;

// calculate derivatives for iphi[int_loopcounter_k]


dbl_par_diphi[int_loopcounter_k]=1;

// calculate derivatives for jphi[int_loopcounter_m]


dbl_par_djphi[int_loopcounter_m]=1;

for (kindex = 0; kindex < nindex; kindex++)


{
k = index[kindex];

// assign values to attributes


dbl_atr_str= Data.d[nstr][k];
dbl_atr_corr_str= Data.d[ncorrstr][k];
dbl_atr_PI= Data.d[nPI][k];
dbl_atr_OCR= Data.d[nOCR][k];
dbl_atr_e= Data.d[ne][k];
dbl_atr_conpre= Data.d[nconpre][k];
dbl_atr_frq= Data.d[nfrq][k];
dbl_atr_N= Data.d[nN][k];
datatype = int(Data.d[nTYPE][k]);

322
// calculation of normalized modulus and damping values
// for given atributes and model parameters

// although the effect of void ratio is not accounted for in this study
// the code is written so that an F(e) term such as the one below
// can be included in the future
// dbl_atr_Fe=0.3 + 0.7 * pow (dbl_atr_e,2);

dbl_atr_Fe=1.0;

dbl_refstr=(dbl_par_phi[1]+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OC
R,dbl_par_phi[3]))

*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4]);

dbl_a=dbl_par_phi[5];

dbl_NG=1.0/(1+pow((dbl_atr_str/dbl_refstr),dbl_a));

dbl_NG_corrstr=1.0/(1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a));

dbl_DMasing=(100.0/dbl_con_pi)*(4*(dbl_atr_corr_str-
dbl_refstr*log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr))

/(pow(dbl_atr_corr_str,2)/(dbl_atr_corr_str+dbl_refstr))-2);

dbl_c1= -1.1143*pow(dbl_a,2)+1.8618*dbl_a+0.2523;

dbl_c2= 0.0805*pow(dbl_a,2)-0.0710*dbl_a-0.0095;

dbl_c3= -0.0005*pow(dbl_a,2)+0.0002*dbl_a+0.0003;

dbl_Dmin=
(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8]))

*pow(dbl_atr_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*log(dbl_atr_fr
q));

dbl_b=dbl_par_phi[11]+dbl_par_phi[12]*log(dbl_atr_N);

323
dbl_D=dbl_Dmin+dbl_b*pow(dbl_NG_corrstr,0.1)*(dbl_c1*dbl_DMasin
g+dbl_c2*pow(dbl_DMasing,2)
+dbl_c3*pow(dbl_DMasing,3));

//"FIRST ORDER DERIVATIVES" CODE FOR iphi[int_loopcounter_k]


//-------------------------------------------------------

dbl_direfstr=dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*dbl_par_di
phi[1]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*dbl_atr_Fe*pow(dbl_atr
_conpre,dbl_par_phi[4])*dbl_par_diphi[2]

+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*log(dbl_
atr_OCR)*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*dbl_par_diphi[3]

+(dbl_par_phi[1]+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_
phi[3]))*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*log(dbl_atr_conpre)*d
bl_par_diphi[4];

dbl_dia=dbl_par_diphi[5];

dbl_diNG=(1/pow((1+pow((dbl_atr_str/dbl_refstr),dbl_a)),2))*pow((dbl_
atr_str/dbl_refstr),dbl_a)*(dbl_a/dbl_refstr)*dbl_direfstr
+(-
1/pow((1+pow((dbl_atr_str/dbl_refstr),dbl_a)),2))*pow((dbl_atr_str/dbl_refstr),db
l_a)*log(dbl_atr_str/dbl_refstr)*dbl_dia;

dbl_diNG_corrstr=(1/pow((1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a)),2)
)*pow((dbl_atr_corr_str/dbl_refstr),dbl_a)*(dbl_a/dbl_refstr)*dbl_direfstr
+(-
1/pow((1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a)),2))*pow((dbl_atr_corr_str/db
l_refstr),dbl_a)*log(dbl_atr_corr_str/dbl_refstr)*dbl_dia;

dbl_dic1=(-2.2286*dbl_a+1.8618)*dbl_dia;

dbl_dic2=( 0.1610*dbl_a-0.0710)*dbl_dia;

dbl_dic3=(-0.0010*dbl_a+0.0002)*dbl_dia;

324
dbl_diDMasing=(-
400)*((log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr)*dbl_atr_corr_str

+2*dbl_refstr*log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr)
-
2*dbl_atr_corr_str)/(pow(dbl_atr_corr_str,2)*dbl_con_pi))
*dbl_direfstr;

dbl_diDmin=pow(dbl_atr_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*lo
g(dbl_atr_frq))*dbl_par_diphi[6]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*pow(dbl_atr_conpre,dbl_par_ph
i[9])*(1+dbl_par_phi[10]*log(dbl_atr_frq))*dbl_par_diphi[7]

+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*log(dbl_atr_OC
R)*pow(dbl_atr_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*log(dbl_atr_frq))*d
bl_par_diphi[8]

+(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])
)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_conpre)*(1+dbl_par_phi[10]*
log(dbl_atr_frq))*dbl_par_diphi[9]

+(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])
)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_frq)*dbl_par_diphi[10];

dbl_dib=dbl_par_diphi[11]+log(dbl_atr_N)*dbl_par_diphi[12];

dbl_diD=dbl_diDmin

+pow(dbl_NG_corrstr,0.1)*dbl_DMasing*(dbl_c1+dbl_c2*dbl_DMasing+dbl_c3
*pow(dbl_DMasing,2))*dbl_dib

+0.1*dbl_b*dbl_DMasing*((dbl_c1+dbl_c2*dbl_DMasing+dbl_c3*pow(dbl_DM
asing,2))/pow(dbl_NG_corrstr,0.9))*dbl_diNG_corrstr

+dbl_b*pow(dbl_NG_corrstr,0.1)*(dbl_c1+2*dbl_c2*dbl_DMasing+3*dbl_c3*p
ow(dbl_DMasing,2))*dbl_diDMasing

+dbl_b*pow(dbl_NG_corrstr,0.1)*dbl_DMasing*dbl_dic1

325
+dbl_b*pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,2)*dbl_dic2

+dbl_b*pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,3)*dbl_dic3;
//-------------------------------------------------------
//"FIRST ORDER DERIVATIVES" CODE FOR jphi[int_loopcounter_m]
//-------------------------------------------------------

dbl_djrefstr=dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*dbl_par_dj
phi[1]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*dbl_atr_Fe*pow(dbl_atr
_conpre,dbl_par_phi[4])*dbl_par_djphi[2]

+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*log(dbl_
atr_OCR)*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*dbl_par_djphi[3]

+(dbl_par_phi[1]+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_
phi[3]))*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*log(dbl_atr_conpre)*d
bl_par_djphi[4];

dbl_dja=dbl_par_djphi[5];

dbl_djNG=(1/pow((1+pow((dbl_atr_str/dbl_refstr),dbl_a)),2))*pow((dbl_
atr_str/dbl_refstr),dbl_a)*(dbl_a/dbl_refstr)*dbl_djrefstr
+(-
1/pow((1+pow((dbl_atr_str/dbl_refstr),dbl_a)),2))*pow((dbl_atr_str/dbl_refstr),db
l_a)*log(dbl_atr_str/dbl_refstr)*dbl_dja;

dbl_djNG_corrstr=(1/pow((1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a)),2)
)*pow((dbl_atr_corr_str/dbl_refstr),dbl_a)*(dbl_a/dbl_refstr)*dbl_djrefstr
+(-
1/pow((1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a)),2))*pow((dbl_atr_corr_str/db
l_refstr),dbl_a)*log(dbl_atr_corr_str/dbl_refstr)*dbl_dja;

dbl_djc1=(-2.2286*dbl_a+1.8618)*dbl_dja;

dbl_djc2=( 0.1610*dbl_a-0.0710)*dbl_dja;

dbl_djc3=(-0.0010*dbl_a+0.0002)*dbl_dja;

326
dbl_djDMasing=(-
400)*((log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr)*dbl_atr_corr_str

+2*dbl_refstr*log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr)
-
2*dbl_atr_corr_str)/(pow(dbl_atr_corr_str,2)*dbl_con_pi))
*dbl_djrefstr;

dbl_djDmin=pow(dbl_atr_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*lo
g(dbl_atr_frq))*dbl_par_djphi[6]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*pow(dbl_atr_conpre,dbl_par_ph
i[9])*(1+dbl_par_phi[10]*log(dbl_atr_frq))*dbl_par_djphi[7]

+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*log(dbl_atr_OC
R)*pow(dbl_atr_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*log(dbl_atr_frq))*d
bl_par_djphi[8]

+(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])
)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_conpre)*(1+dbl_par_phi[10]*
log(dbl_atr_frq))*dbl_par_djphi[9]

+(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])
)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_frq)*dbl_par_djphi[10];

dbl_djb=dbl_par_djphi[11]+log(dbl_atr_N)*dbl_par_djphi[12];

dbl_djD=dbl_djDmin

+pow(dbl_NG_corrstr,0.1)*dbl_DMasing*(dbl_c1+dbl_c2*dbl_DMasing+dbl_c3
*pow(dbl_DMasing,2))*dbl_djb

+0.1*dbl_b*dbl_DMasing*((dbl_c1+dbl_c2*dbl_DMasing+dbl_c3*pow(dbl_DM
asing,2))/pow(dbl_NG_corrstr,0.9))*dbl_djNG_corrstr

+dbl_b*pow(dbl_NG_corrstr,0.1)*(dbl_c1+2*dbl_c2*dbl_DMasing+3*dbl_c3*p
ow(dbl_DMasing,2))*dbl_djDMasing

+dbl_b*pow(dbl_NG_corrstr,0.1)*dbl_DMasing*dbl_djc1

327
+dbl_b*pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,2)*dbl_djc2

+dbl_b*pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,3)*dbl_djc3;
//-------------------------------------------------------
//"SECOND ORDER DERIVATIVES" CODE FOR
iphi[int_loopcounter_k],jphi[int_loopcounter_m]
// Requires first order derivatives with respect to i and j
//-------------------------------------------------------

dbl_d2refstr=dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*log(dbl_atr
_conpre)*dbl_par_diphi[1]*dbl_par_djphi[4]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*log(dbl_atr_OCR)*dbl_atr_Fe*p
ow(dbl_atr_conpre,dbl_par_phi[4])*dbl_par_diphi[2]*dbl_par_djphi[3]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*dbl_atr_Fe*pow(dbl_atr_conpre
,dbl_par_phi[4])*log(dbl_atr_conpre)*dbl_par_diphi[2]*dbl_par_djphi[4]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*log(dbl_atr_OCR)*dbl_atr_Fe*p
ow(dbl_atr_conpre,dbl_par_phi[4])*dbl_par_diphi[3]*dbl_par_djphi[2]

+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*pow(log(dbl_atr
_OCR),2)*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*dbl_par_diphi[3]*db
l_par_djphi[3]

+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*log(dbl_atr_OC
R)*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*log(dbl_atr_conpre)*dbl_pa
r_diphi[3]*dbl_par_djphi[4]

+dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*log(dbl_atr_conpre)*dbl_par_
diphi[4]*dbl_par_djphi[1]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*dbl_atr_Fe*pow(dbl_atr_conpre
,dbl_par_phi[4])*log(dbl_atr_conpre)*dbl_par_diphi[4]*dbl_par_djphi[2]

+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])*log(dbl_atr_OC
R)*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*log(dbl_atr_conpre)*dbl_pa
r_diphi[4]*dbl_par_djphi[3]

+(dbl_par_phi[1]+dbl_par_phi[2]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[3])

328
)*dbl_atr_Fe*pow(dbl_atr_conpre,dbl_par_phi[4])*pow(log(dbl_atr_conpre),2)*d
bl_par_diphi[4]*dbl_par_djphi[4];

dbl_d2a=0;

dbl_d2NG=dbl_a*((pow((dbl_atr_str/dbl_refstr),(2*dbl_a))*dbl_a
-
pow((dbl_atr_str/dbl_refstr),dbl_a)*dbl_a
-
pow((dbl_atr_str/dbl_refstr),dbl_a)
-
pow((dbl_atr_str/dbl_refstr),(2*dbl_a)))

/(pow(dbl_refstr,2)*pow((1+pow((dbl_atr_str/dbl_refstr),dbl_a)),3)))*dbl_
direfstr*dbl_djrefstr
+((-
2*pow(pow(dbl_atr_str,2),dbl_a)*pow((1/pow(dbl_refstr,2)),dbl_a)*dbl_a*log(db
l_atr_str)

+2*pow(pow(dbl_atr_str,2),dbl_a)*pow((1/pow(dbl_refstr,2)),dbl_a)*dbl_a*log(
dbl_refstr)

+pow(dbl_atr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)*dbl_a*log(dbl_atr_str)
-
pow(dbl_atr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)*dbl_a*log(dbl_refstr)

+pow(dbl_atr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a))*dbl_a*log(dbl_atr_str
)
-
pow(dbl_atr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a))*dbl_a*log(dbl_refstr)

+pow(dbl_atr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)

+pow(dbl_atr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a)))

/(dbl_refstr*pow((1+pow(dbl_atr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)),3)))*dbl_
direfstr*dbl_dja
+((-
2*pow(pow(dbl_atr_str,2),dbl_a)*pow((1/pow(dbl_refstr,2)),dbl_a)*dbl_a*log(db
l_atr_str)

329
+2*pow(pow(dbl_atr_str,2),dbl_a)*pow((1/pow(dbl_refstr,2)),dbl_a)*dbl_a*log(
dbl_refstr)

+pow(dbl_atr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)*dbl_a*log(dbl_atr_str)
-
pow(dbl_atr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)*dbl_a*log(dbl_refstr)

+pow(dbl_atr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a))*dbl_a*log(dbl_atr_str
)
-
pow(dbl_atr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a))*dbl_a*log(dbl_refstr)

+pow(dbl_atr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)

+pow(dbl_atr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a)))

/(dbl_refstr*pow((1+pow(dbl_atr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)),3)))*dbl_
dia*dbl_djrefstr
+(-
pow(log(dbl_atr_str/dbl_refstr),2)*(-pow((dbl_atr_str/dbl_refstr),(2*dbl_a))

+pow((dbl_atr_str/dbl_refstr),dbl_a))

/pow((1+pow((dbl_atr_str/dbl_refstr),dbl_a)),3))*dbl_dia*dbl_dja

+((1/pow((1+pow((dbl_atr_str/dbl_refstr),dbl_a)),2))*pow((dbl_atr_str/db
l_refstr),dbl_a)*(dbl_a/dbl_refstr)*dbl_d2refstr
+(-
1/pow((1+pow((dbl_atr_str/dbl_refstr),dbl_a)),2))*pow((dbl_atr_str/dbl_refstr),db
l_a)*log(dbl_atr_str/dbl_refstr)*dbl_d2a);

dbl_d2NG_corrstr=dbl_a*((pow((dbl_atr_corr_str/dbl_refstr),(2*dbl_a))*
dbl_a
-
pow((dbl_atr_corr_str/dbl_refstr),dbl_a)*dbl_a
-
pow((dbl_atr_corr_str/dbl_refstr),dbl_a)
-
pow((dbl_atr_corr_str/dbl_refstr),(2*dbl_a)))

330
/(pow(dbl_refstr,2)*pow((1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a)),3)))
*dbl_direfstr*dbl_djrefstr
+((-
2*pow(pow(dbl_atr_corr_str,2),dbl_a)*pow((1/pow(dbl_refstr,2)),dbl_a)*dbl_a*l
og(dbl_atr_corr_str)

+2*pow(pow(dbl_atr_corr_str,2),dbl_a)*pow((1/pow(dbl_refstr,2)),dbl_a)*dbl_a
*log(dbl_refstr)

+pow(dbl_atr_corr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)*dbl_a*log(dbl_atr_corr_
str)
-
pow(dbl_atr_corr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)*dbl_a*log(dbl_refstr)

+pow(dbl_atr_corr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a))*dbl_a*log(dbl_a
tr_corr_str)
-
pow(dbl_atr_corr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a))*dbl_a*log(dbl_re
fstr)

+pow(dbl_atr_corr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)

+pow(dbl_atr_corr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a)))

/(dbl_refstr*pow((1+pow(dbl_atr_corr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)),3)))
*dbl_direfstr*dbl_dja
+((-
2*pow(pow(dbl_atr_corr_str,2),dbl_a)*pow((1/pow(dbl_refstr,2)),dbl_a)*dbl_a*l
og(dbl_atr_corr_str)

+2*pow(pow(dbl_atr_corr_str,2),dbl_a)*pow((1/pow(dbl_refstr,2)),dbl_a)*dbl_a
*log(dbl_refstr)

+pow(dbl_atr_corr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)*dbl_a*log(dbl_atr_corr_
str)
-
pow(dbl_atr_corr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)*dbl_a*log(dbl_refstr)

+pow(dbl_atr_corr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a))*dbl_a*log(dbl_a
tr_corr_str)

331
-
pow(dbl_atr_corr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a))*dbl_a*log(dbl_re
fstr)

+pow(dbl_atr_corr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)

+pow(dbl_atr_corr_str,(2*dbl_a))*pow((1/dbl_refstr),(2*dbl_a)))

/(dbl_refstr*pow((1+pow(dbl_atr_corr_str,dbl_a)*pow((1/dbl_refstr),dbl_a)),3)))
*dbl_dia*dbl_djrefstr
+(-
pow(log(dbl_atr_corr_str/dbl_refstr),2)*(-
pow((dbl_atr_corr_str/dbl_refstr),(2*dbl_a))

+pow((dbl_atr_corr_str/dbl_refstr),dbl_a))

/pow((1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a)),3))*dbl_dia*dbl_dja

+((1/pow((1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a)),2))*pow((dbl_atr_
corr_str/dbl_refstr),dbl_a)*(dbl_a/dbl_refstr)*dbl_d2refstr
+(-
1/pow((1+pow((dbl_atr_corr_str/dbl_refstr),dbl_a)),2))*pow((dbl_atr_corr_str/db
l_refstr),dbl_a)*log(dbl_atr_corr_str/dbl_refstr)*dbl_d2a);

dbl_d2DMasing=400*((pow(dbl_atr_corr_str,2)
-
2*dbl_refstr*log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr)*dbl_atr_corr_str
-
2*log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr)*pow(dbl_refstr,2)

+2*dbl_atr_corr_str*dbl_refstr)

/(dbl_con_pi*dbl_refstr*(dbl_atr_corr_str+dbl_refstr)*pow(dbl_atr_corr_s
tr,2)))

*dbl_direfstr*dbl_djrefstr
+(-
400)*((log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr)*dbl_atr_corr_str

+2*dbl_refstr*log((dbl_atr_corr_str+dbl_refstr)/dbl_refstr)

332
-
2*dbl_atr_corr_str)/(pow(dbl_atr_corr_str,2)*dbl_con_pi))

*dbl_d2refstr;

dbl_d2c1=-2.2286*dbl_dia*dbl_dja+(-
2.2286*dbl_a+1.8618)*dbl_d2a;

dbl_d2c2= 0.1610*dbl_dia*dbl_dja+(
0.1610*dbl_a-0.0710)*dbl_d2a;

dbl_d2c3=-0.0010*dbl_dia*dbl_dja+(-
0.0010*dbl_a+0.0002)*dbl_d2a;

dbl_d2Dmin=pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_conpre)*(
1+dbl_par_phi[10]*log(dbl_atr_frq))*dbl_par_diphi[6]*dbl_par_djphi[9]

+pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_frq)*dbl_par_diphi[6]*dbl_pa
r_djphi[10]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*log(dbl_atr_OCR)*pow(dbl_atr
_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*log(dbl_atr_frq))*dbl_par_diphi[7]
*dbl_par_djphi[8]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*pow(dbl_atr_conpre,dbl_par_ph
i[9])*log(dbl_atr_conpre)*(1+dbl_par_phi[10]*log(dbl_atr_frq))*dbl_par_diphi[7
]*dbl_par_djphi[9]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*pow(dbl_atr_conpre,dbl_par_ph
i[9])*log(dbl_atr_frq)*dbl_par_diphi[7]*dbl_par_djphi[10]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*log(dbl_atr_OCR)*pow(dbl_atr
_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*log(dbl_atr_frq))*dbl_par_diphi[8]
*dbl_par_djphi[7]

+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*pow(log(dbl_atr
_OCR),2)*pow(dbl_atr_conpre,dbl_par_phi[9])*(1+dbl_par_phi[10]*log(dbl_atr
_frq))*dbl_par_diphi[8]*dbl_par_djphi[8]

+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*log(dbl_atr_OC

333
R)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_conpre)*(1+dbl_par_phi[10
]*log(dbl_atr_frq))*dbl_par_diphi[8]*dbl_par_djphi[9]

+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*log(dbl_atr_OC
R)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_frq)*dbl_par_diphi[8]*dbl_
par_djphi[10]

+pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_conpre)*(1+dbl_par_phi[10]*l
og(dbl_atr_frq))*dbl_par_diphi[9]*dbl_par_djphi[6]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*pow(dbl_atr_conpre,dbl_par_ph
i[9])*log(dbl_atr_conpre)*(1+dbl_par_phi[10]*log(dbl_atr_frq))*dbl_par_diphi[9
]*dbl_par_djphi[7]

+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*log(dbl_atr_OC
R)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_conpre)*(1+dbl_par_phi[10
]*log(dbl_atr_frq))*dbl_par_diphi[9]*dbl_par_djphi[8]

+(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])
)*pow(dbl_atr_conpre,dbl_par_phi[9])*pow(log(dbl_atr_conpre),2)*(1+dbl_par_
phi[10]*log(dbl_atr_frq))*dbl_par_diphi[9]*dbl_par_djphi[9]

+(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])
)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_conpre)*log(dbl_atr_frq)*dbl
_par_diphi[9]*dbl_par_djphi[10]

+pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_frq)*dbl_par_diphi[10]*dbl_p
ar_djphi[6]

+dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*pow(dbl_atr_conpre,dbl_par_ph
i[9])*log(dbl_atr_frq)*dbl_par_diphi[10]*dbl_par_djphi[7]

+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])*log(dbl_atr_OC
R)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_frq)*dbl_par_diphi[10]*dbl
_par_djphi[8]

+(dbl_par_phi[6]+dbl_par_phi[7]*dbl_atr_PI*pow(dbl_atr_OCR,dbl_par_phi[8])
)*pow(dbl_atr_conpre,dbl_par_phi[9])*log(dbl_atr_conpre)*log(dbl_atr_frq)*dbl
_par_diphi[10]*dbl_par_djphi[9];

dbl_d2b=0;

334
dbl_d2D=0.1*dbl_DMasing*((dbl_c1+dbl_c2*dbl_DMasing+dbl_c3*pow
(dbl_DMasing,2))/pow(dbl_NG_corrstr,0.9))*dbl_dib*dbl_djNG_corrstr

+pow(dbl_NG_corrstr,0.1)*(dbl_c1+2*dbl_c2*dbl_DMasing+3*dbl_c3*pow(dbl
_DMasing,2))*dbl_dib*dbl_djDMasing

+pow(dbl_NG_corrstr,0.1)*dbl_DMasing*dbl_dib*dbl_djc1

+pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,2)*dbl_dib*dbl_djc2

+pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,3)*dbl_dib*dbl_djc3

+0.1*dbl_DMasing*((dbl_c1+dbl_c2*dbl_DMasing+dbl_c3*pow(dbl_D
Masing,2))/pow(dbl_NG_corrstr,0.9))*dbl_diNG_corrstr*dbl_djb
+(-
0.09)*dbl_b*dbl_DMasing*((dbl_c1+dbl_c2*dbl_DMasing+dbl_c3*pow(dbl_D
Masing,2))/pow(dbl_NG_corrstr,1.9))*dbl_diNG_corrstr*dbl_djNG_corrstr

+0.1*(dbl_b/pow(dbl_NG_corrstr,0.9))*(dbl_c1+2*dbl_c2*dbl_DMasing
+3*dbl_c3*pow(dbl_DMasing,2))*dbl_diNG_corrstr*dbl_djDMasing

+0.1*(dbl_b/pow(dbl_NG_corrstr,0.9))*dbl_DMasing*dbl_diNG_corrstr*
dbl_djc1

+0.1*(dbl_b/pow(dbl_NG_corrstr,0.9))*pow(dbl_DMasing,2)*dbl_diNG_
corrstr*dbl_djc2

+0.1*(dbl_b/pow(dbl_NG_corrstr,0.9))*pow(dbl_DMasing,3)*dbl_diNG_
corrstr*dbl_djc3

+pow(dbl_NG_corrstr,0.1)*(dbl_c1+2*dbl_c2*dbl_DMasing+3*dbl_c3*p
ow(dbl_DMasing,2))*dbl_diDMasing*dbl_djb

+0.1*(dbl_b/pow(dbl_NG_corrstr,0.9))*(dbl_c1+2*dbl_c2*dbl_DMasing
+3*dbl_c3*pow(dbl_DMasing,2))*dbl_diDMasing*dbl_djNG_corrstr

+2*dbl_b*pow(dbl_NG_corrstr,0.1)*(dbl_c2+3*dbl_c3*dbl_DMasing)*d
bl_diDMasing*dbl_djDMasing

+dbl_b*pow(dbl_NG_corrstr,0.1)*dbl_diDMasing*dbl_djc1

335
+2*dbl_b*pow(dbl_NG_corrstr,0.1)*dbl_DMasing*dbl_diDMasing*dbl_
djc2

+3*dbl_b*pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,2)*dbl_diDMasin
g*dbl_djc3

+pow(dbl_NG_corrstr,0.1)*dbl_DMasing*dbl_dic1*dbl_djb

+0.1*(dbl_b/pow(dbl_NG_corrstr,0.9))*dbl_DMasing*dbl_dic1*dbl_djN
G_corrstr

+dbl_b*pow(dbl_NG_corrstr,0.1)*dbl_dic1*dbl_djDMasing

+pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,2)*dbl_dic2*dbl_djb

+0.1*(dbl_b/pow(dbl_NG_corrstr,0.9))*pow(dbl_DMasing,2)*dbl_dic2*d
bl_djNG_corrstr

+2*dbl_b*pow(dbl_NG_corrstr,0.1)*dbl_DMasing*dbl_dic2*dbl_djDMa
sing

+pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,3)*dbl_dic3*dbl_djb

+0.1*(dbl_b/pow(dbl_NG_corrstr,0.9))*pow(dbl_DMasing,3)*dbl_dic3*d
bl_djNG_corrstr

+3*dbl_b*pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,2)*dbl_dic3*dbl_
djDMasing
+dbl_d2Dmin

+pow(dbl_NG_corrstr,0.1)*dbl_DMasing*(dbl_c1+dbl_c2*dbl_DMasing
+dbl_c3*pow(dbl_DMasing,2))*dbl_d2b

+0.1*dbl_b*dbl_DMasing*((dbl_c1+dbl_c2*dbl_DMasing+dbl_c3*pow(
dbl_DMasing,2))/pow(dbl_NG_corrstr,0.9))*dbl_d2NG_corrstr

+dbl_b*pow(dbl_NG_corrstr,0.1)*(dbl_c1+2*dbl_c2*dbl_DMasing+3*db
l_c3*pow(dbl_DMasing,2))*dbl_d2DMasing

+dbl_b*pow(dbl_NG_corrstr,0.1)*dbl_DMasing*dbl_d2c1

336
+dbl_b*pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,2)*dbl_d2c2

+dbl_b*pow(dbl_NG_corrstr,0.1)*pow(dbl_DMasing,3)*dbl_d2c3;
//-------------------------------------------------------

if (datatype == 0)
{
dYMeani[kindex] = scalar*dbl_diNG;
}
else
{
dYMeani[kindex] = scalar*dbl_diD;
}

if (datatype == 0)
{
dYMeanj[kindex] = scalar*dbl_djNG;
}
else
{
dYMeanj[kindex] = scalar*dbl_djD;
}

if (datatype == 0)
{
d2YMeanij[kindex] = scalar*dbl_d2NG;
}
else
{
d2YMeanij[kindex] = scalar*dbl_d2D;
}

} // for (kindex = 0; kindex < nindex; kindex++)


}

337
APPENDIX D

FILE USED IN ESTIMATING COVARIANCE STRUCTURE

FOR

FIRST ORDER SECOND MOMENT

BAYESIAN ANALYSIS

OF

RESONANT COLUMN

AND

TORSIONAL SHEAR

TEST RESULTS

338
// RCTSYCOV.cpp : Covariance Structure for RCTS Data
//
#include "stdafx.h"
#include <afxwin.h>
#include <iostream.h>
#include <fstream.h>
#include <math.h>
#include <time.h>
#include <direct.h>
#include "machh.h"
#include "compareh.h"
#include "dblash.h"
#include "_arrayh.h"
#include "_array2h.h"
#include "_array3h.h"
#include "matrixh.h"
#include "smatrixh.h"
#include "covmatrixh.h"
#include "gmatrixh.h"
#include "dblash.h"
#include "goldenh.h"
#include "rqph.h"
#include "Datah.h"
#include "NormalLikeh.h"
#include "Modelh.h"

void ModelStructure::CalculateYCOVC(DataStructure &Data, double *x, darray


&YMean, CovMatrix &YCOV,
iarray &index)
{
// Fill and Invert Conditional COV Matrix
int ka,kb,n,kaindex,kbindex;
double Ga,Gb,Da,Db;
int datatypea;

n = index.n;

Ga = exp(x[istdGa]);
Gb = exp(x[istdGb]);
Da = exp(x[istdDa]);
Db = exp(x[istdDb]);

339
for (kaindex = 0; kaindex < n; kaindex++)
{
ka = index[kaindex];
datatypea = int(Data.d[nTYPE][ka]);

if (datatypea == 0)
{
YCOV.G.xptr[kaindex] = scalar*(Ga + pow(((0.25/Gb)-
(pow((YMean[kaindex]-0.5),2)/Gb)),0.5));
}
else
{
YCOV.G.xptr[kaindex] = scalar*(Da +
Db*pow(YMean[kaindex],0.5));
}

for (kbindex = 0; kbindex < kaindex; kbindex++)


{
kb = index[kbindex];
YCOV.R.a[kaindex][kbindex] =
CalculateYrhoab(Data,x,ka,kb);
}
YCOV.R.a[kaindex][kaindex] = 1.0;
}
YCOV.Decompose(0);
}

void ModelStructure::CalculatedYCOVCiMM(int iv, DataStructure &Data,


double *x, darray &YMean,
darray &dYMeani, CovMatrix
&YCOV, smatrixsolve &dYCOVi, iarray &index)
{
int ka,kb,n,kaindex,kbindex;
double Ga,Gb,Da,Db;
double dGa,dGai,dGb,dGbi;
double dDa,dDai,dDb,dDbi;
double dYCOVipartsGka,dYCOVipartsGkb;
double dYCOVipartsRkakb;
int datatypea,datatypeb;

darray tau;
tau.construct(ntheta);

340
n = index.n;

Ga = exp(x[istdGa]);
dGa = Ga;
dGai = double(iv == istdGa)*dGa;
Gb = exp(x[istdGb]);
dGb = Gb;
dGbi = double(iv == istdGb)*dGb;
Da = exp(x[istdDa]);
dDa = Da;
dDai = double(iv == istdDa)*dDa;
Db = exp(x[istdDb]);
dDb = Db;
dDbi = double(iv == istdDb)*dDb;

for (kaindex = 0; kaindex < n; kaindex++)


{
ka = index[kaindex];
datatypea = int(Data.d[nTYPE][ka]);

if (datatypea == 0)
{
dYCOVipartsGka = scalar*( dGai
+
dGbi*(-1/(2*pow(((0.25/Gb)-(pow((YMean[kaindex]-0.5),2)/Gb)),0.5)))

*((0.25/Gb)-(pow((YMean[kaindex]-0.5),2)/Gb))*(1/Gb)
+
dYMeani[kaindex]*(-1/pow(((0.25/Gb)-(pow((YMean[kaindex]-
0.5),2)/Gb)),0.5))

*(YMean[kaindex]-0.5)/Gb);
}
else
{
dYCOVipartsGka = scalar*( dDai
+
dDbi*pow(YMean[kaindex],0.5)
+
dYMeani[kaindex]*0.5*Db/pow(YMean[kaindex],0.5));
}

341
for (kbindex = 0; kbindex < kaindex; kbindex++)
{
kb = index[kbindex];
datatypeb = int(Data.d[nTYPE][kb]);

if (datatypeb == 0)
{
dYCOVipartsGkb = scalar*( dGai
+
dGbi*(-1/(2*pow(((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb)),0.5)))

*((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb))*(1/Gb)
+
dYMeani[kbindex]*(-1/pow(((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb)),0.5))

*(YMean[kbindex]-0.5)/Gb);
}
else
{
dYCOVipartsGkb = scalar*( dDai
+
dDbi*pow(YMean[kbindex],0.5)
+
dYMeani[kbindex]*0.5*Db/pow(YMean[kbindex],0.5));
}

tau[0] = fabs(log(Data.d[nstr][ka])-log(Data.d[nstr][kb]));
tau[1] = double (Data.d[nTYPE][ka] !=
Data.d[nTYPE][kb]);
tau[2] = double (Data.d[nspecimen][ka] !=
Data.d[nspecimen][kb]);
tau[3] = double (Data.d[ntest][ka] != Data.d[ntest][kb]);
tau[4] = double (Data.d[npressure][ka] !=
Data.d[npressure][kb]);

dYCOVipartsRkakb = dYCOVrhoi(iv,tau.xptr,x);
dYCOVi.a[kaindex][kbindex] =
dYCOVipartsGka*YCOV.R.a[kaindex][kbindex]*YCOV.G.xptr[kbindex]

342
+
YCOV.G.xptr[kaindex]*dYCOVipartsRkakb*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*YCOV.R.a[kaindex][kbindex]*dYCOVipartsGkb;
}

// kbindex = kaindex
kb = index[kbindex];
datatypeb = int(Data.d[nTYPE][kb]);

if (datatypeb == 0)
{
dYCOVipartsGkb = scalar*( dGai
+
dGbi*(-1/(2*pow(((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb)),0.5)))

*((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb))*(1/Gb)
+
dYMeani[kbindex]*(-1/pow(((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb)),0.5))

*(YMean[kbindex]-0.5)/Gb);
}
else
{
dYCOVipartsGkb = scalar*( dDai
+
dDbi*pow(YMean[kbindex],0.5)
+
dYMeani[kbindex]*0.5*Db/pow(YMean[kbindex],0.5));
}

dYCOVipartsRkakb = 0.0;
dYCOVi.a[kaindex][kbindex] =
dYCOVipartsGka*YCOV.R.a[kaindex][kbindex]*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*dYCOVipartsRkakb*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*YCOV.R.a[kaindex][kbindex]*dYCOVipartsGkb;
}
}

343
void ModelStructure::Calculated2YCOVCijMM(int iv, int jv, DataStructure
&Data, double *x, darray &YMean,
darray &dYMeani, darray
&dYMeanj, darray &d2YMeanij, CovMatrix &YCOV,
smatrixsolve &dYCOVi,
smatrixsolve &dYCOVj, smatrixsolve &d2YCOVij, iarray &index)
{
int ka,kb,n,kaindex,kbindex;
double Ga,Gb,Da,Db;
double dGa,dGai,dGaj,dGb,dGbi,dGbj;
double dDa,dDai,dDaj,dDb,dDbi,dDbj;
double d2Ga,d2Gaij,d2Gb,d2Gbij;
double d2Da,d2Daij,d2Db,d2Dbij;
double
dYCOVipartsGka,dYCOVjpartsGka,dYCOVipartsGkb,dYCOVjpartsGkb;
double d2YCOVijGka,d2YCOVijGkb;
double dYCOVipartsRkakb,dYCOVjpartsRkakb;
double d2YCOVijRab;
int datatypea,datatypeb;

darray tau;
tau.construct(ntheta);

n = index.n;
darray d2YCOVijG(n);

Ga = exp(x[istdGa]);
dGa = Ga;
d2Ga = Ga;
dGai = double(iv == istdGa)*dGa;
dGaj = double(jv == istdGa)*dGa;
d2Gaij = double(iv == istdGa)*double(jv == istdGa)*d2Ga;
Gb = exp(x[istdGb]);
dGb = Gb;
d2Gb = Gb;
dGbi = double(iv == istdGb)*dGb;
dGbj = double(jv == istdGb)*dGb;
d2Gbij = double(iv == istdGb)*double(jv == istdGb)*d2Gb;
Da = exp(x[istdDa]);
dDa = Da;
d2Da = Da;
dDai = double(iv == istdDa)*dDa;

344
dDaj = double(jv == istdDa)*dDa;
d2Daij = double(iv == istdDa)*double(jv == istdDa)*d2Da;
Db = exp(x[istdDb]);
dDb = Db;
d2Db = Db;
dDbi = double(iv == istdDb)*dDb;
dDbj = double(jv == istdDb)*dDb;
d2Dbij = double(iv == istdDb)*double(jv == istdDb)*d2Db;

for (kaindex = 0; kaindex < n; kaindex++)


{
ka = index[kaindex];
datatypea = int(Data.d[nTYPE][ka]);

if (datatypea == 0)
{
dYCOVipartsGka = scalar*( dGai
+
dGbi*(-1/(2*pow(((0.25/Gb)-(pow((YMean[kaindex]-0.5),2)/Gb)),0.5)))

*((0.25/Gb)-(pow((YMean[kaindex]-0.5),2)/Gb))*(1/Gb)
+
dYMeani[kaindex]*(-1/pow(((0.25/Gb)-(pow((YMean[kaindex]-
0.5),2)/Gb)),0.5))

*(YMean[kaindex]-0.5)/Gb);
dYCOVjpartsGka = scalar*( dGaj
+
dGbj*(-1/(2*pow(((0.25/Gb)-(pow((YMean[kaindex]-0.5),2)/Gb)),0.5)))

*((0.25/Gb)-(pow((YMean[kaindex]-0.5),2)/Gb))*(1/Gb)
+
dYMeanj[kaindex]*(-1/pow(((0.25/Gb)-(pow((YMean[kaindex]-
0.5),2)/Gb)),0.5))

*(YMean[kaindex]-0.5)/Gb);
d2YCOVijGka = scalar*( d2Gaij
+
d2Gbij *(1/Gb)* ( (-1/(4*pow(((0.25/Gb)-(pow((YMean[kaindex]-
0.5),2)/Gb)),1.5)))*pow(((0.25/Gb)-(pow((YMean[kaindex]-0.5),2)/Gb)),2) +
(0.5 * pow(((0.25/Gb)-(pow((YMean[kaindex]-0.5),2)/Gb)),0.5) ) )

345
+
dGbi*dYMeanj[kaindex] *(1/Gb)* ( (-1/(2*pow(((0.25/Gb)-
(pow((YMean[kaindex]-0.5),2)/Gb)),1.5)))*((0.25/Gb)-(pow((YMean[kaindex]-
0.5),2)/Gb))*((YMean[kaindex]-0.5)/Gb) + (1/(pow(((0.25/Gb)-
(pow((YMean[kaindex]-0.5),2)/Gb)),0.5)))*((YMean[kaindex]-0.5)/Gb) )
+
dYMeani[kaindex]*dGbj *(1/Gb)* ( (-1/(2*pow(((0.25/Gb)-
(pow((YMean[kaindex]-0.5),2)/Gb)),1.5)))*((0.25/Gb)-(pow((YMean[kaindex]-
0.5),2)/Gb))*((YMean[kaindex]-0.5)/Gb) + (1/(pow(((0.25/Gb)-
(pow((YMean[kaindex]-0.5),2)/Gb)),0.5)))*((YMean[kaindex]-0.5)/Gb) )
+
dYMeani[kaindex]*dYMeanj[kaindex] * ( ((-1/(pow(((0.25/Gb)-
(pow((YMean[kaindex]-0.5),2)/Gb)),1.5)))*( pow((YMean[kaindex]-
0.5),2)/pow(Gb,2) )) - (1/( pow(((0.25/Gb)-(pow((YMean[kaindex]-
0.5),2)/Gb)),0.5)*Gb)) )
+
d2YMeanij[kaindex] * (-1/pow(((0.25/Gb)-(pow((YMean[kaindex]-
0.5),2)/Gb)),0.5))

*(YMean[kaindex]-0.5)/Gb );
}
else
{
dYCOVipartsGka = scalar*( dDai
+
dDbi*pow(YMean[kaindex],0.5)
+
dYMeani[kaindex]*0.5*Db/pow(YMean[kaindex],0.5));
dYCOVjpartsGka = scalar*( dDaj
+
dDbj*pow(YMean[kaindex],0.5)
+
dYMeanj[kaindex]*0.5*Db/pow(YMean[kaindex],0.5));
d2YCOVijGka = scalar*( d2Daij
+
d2Dbij*pow(YMean[kaindex],0.5)
+
dDbi*dYMeanj[kaindex] * (1/(2*pow(YMean[kaindex],0.5)))
+
dYMeani[kaindex]*dDbj * (1/(2*pow(YMean[kaindex],0.5)))
+
dYMeani[kaindex]*dYMeanj[kaindex]*(-0.25)*Db/pow(YMean[kaindex],1.5)

346
+
d2YMeanij[kaindex] * 0.5*Db/pow(YMean[kaindex],0.5) );
}

for (kbindex = 0; kbindex < kaindex; kbindex++)


{
kb = index[kbindex];
datatypeb = int(Data.d[nTYPE][kb]);

if (datatypeb == 0)
{
dYCOVipartsGkb = scalar*( dGai

+ dGbi*(-1/(2*pow(((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb)),0.5)))

*((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb))*(1/Gb)

+ dYMeani[kbindex]*(-1/pow(((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb)),0.5))

*(YMean[kbindex]-0.5)/Gb);
dYCOVjpartsGkb = scalar*( dGaj

+ dGbj*(-1/(2*pow(((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb)),0.5)))

*((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb))*(1/Gb)

+ dYMeanj[kbindex]*(-1/pow(((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb)),0.5))

*(YMean[kbindex]-0.5)/Gb);
d2YCOVijGkb = scalar*( d2Gaij

+ d2Gbij *(1/Gb)* ( (-1/(4*pow(((0.25/Gb)-(pow((YMean[kbindex]-


0.5),2)/Gb)),1.5)))*pow(((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb)),2) +
(0.5 * pow(((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb)),0.5) ) )

+ dGbi*dYMeanj[kbindex] *(1/Gb)* ( (-1/(2*pow(((0.25/Gb)-


(pow((YMean[kbindex]-0.5),2)/Gb)),1.5)))*((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb))*((YMean[kbindex]-0.5)/Gb) + (1/(pow(((0.25/Gb)-
(pow((YMean[kbindex]-0.5),2)/Gb)),0.5)))*((YMean[kbindex]-0.5)/Gb) )

347
+ dYMeani[kbindex]*dGbj *(1/Gb)* ( (-1/(2*pow(((0.25/Gb)-
(pow((YMean[kbindex]-0.5),2)/Gb)),1.5)))*((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb))*((YMean[kbindex]-0.5)/Gb) + (1/(pow(((0.25/Gb)-
(pow((YMean[kbindex]-0.5),2)/Gb)),0.5)))*((YMean[kbindex]-0.5)/Gb) )

+ dYMeani[kbindex]*dYMeanj[kbindex] * ( ((-1/(pow(((0.25/Gb)-
(pow((YMean[kbindex]-0.5),2)/Gb)),1.5)))*( pow((YMean[kbindex]-
0.5),2)/pow(Gb,2) )) - (1/( pow(((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb)),0.5)*Gb)) )

+ d2YMeanij[kbindex] * (-1/pow(((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb)),0.5))

*(YMean[kbindex]-0.5)/Gb );
}
else
{
dYCOVipartsGkb = scalar*( dDai

+ dDbi*pow(YMean[kbindex],0.5)

+ dYMeani[kbindex]*0.5*Db/pow(YMean[kbindex],0.5));
dYCOVjpartsGkb = scalar*( dDaj

+ dDbj*pow(YMean[kbindex],0.5)

+ dYMeanj[kbindex]*0.5*Db/pow(YMean[kbindex],0.5));
d2YCOVijGkb = scalar*( d2Daij

+ d2Dbij*pow(YMean[kbindex],0.5)

+ dDbi*dYMeanj[kbindex] * (1/(2*pow(YMean[kbindex],0.5)))

+ dYMeani[kbindex]*dDbj * (1/(2*pow(YMean[kbindex],0.5)))

+ dYMeani[kbindex]*dYMeanj[kbindex]*(-
0.25)*Db/pow(YMean[kbindex],1.5)

+ d2YMeanij[kbindex] * 0.5*Db/pow(YMean[kbindex],0.5) );
}

348
tau[0] = fabs(log(Data.d[nstr][ka])-log(Data.d[nstr][kb]));
tau[1] = double (Data.d[nTYPE][ka] !=
Data.d[nTYPE][kb]);
tau[2] = double (Data.d[nspecimen][ka] !=
Data.d[nspecimen][kb]);
tau[3] = double (Data.d[ntest][ka] != Data.d[ntest][kb]);
tau[4] = double (Data.d[npressure][ka] !=
Data.d[npressure][kb]);

dYCOVipartsRkakb = dYCOVrhoi(iv,tau.xptr,x);
dYCOVjpartsRkakb = dYCOVrhoi(jv,tau.xptr,x);
dYCOVi.a[kaindex][kbindex] =
dYCOVipartsGka*YCOV.R.a[kaindex][kbindex]*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*dYCOVipartsRkakb*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*YCOV.R.a[kaindex][kbindex]*dYCOVipartsGkb;
dYCOVj.a[kaindex][kbindex] =
dYCOVjpartsGka*YCOV.R.a[kaindex][kbindex]*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*dYCOVjpartsRkakb*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*YCOV.R.a[kaindex][kbindex]*dYCOVjpartsGkb;
d2YCOVijRab = d2YCOVrhoij(iv,jv,tau.xptr,x);
d2YCOVij.a[kaindex][kbindex] =
d2YCOVijGka*YCOV.R.a[kaindex][kbindex]*YCOV.G.xptr[kbindex]
+
dYCOVipartsGka*dYCOVjpartsRkakb*YCOV.G.xptr[kbindex]
+
dYCOVipartsGka*YCOV.R.a[kaindex][kbindex]*dYCOVjpartsGkb
+
dYCOVjpartsGka*dYCOVipartsRkakb*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*d2YCOVijRab*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*dYCOVipartsRkakb*dYCOVjpartsGkb
+
dYCOVjpartsGka*YCOV.R.a[kaindex][kbindex]*dYCOVipartsGkb
+
YCOV.G.xptr[kaindex]*dYCOVjpartsRkakb*dYCOVipartsGkb
+
YCOV.G.xptr[kaindex]*YCOV.R.a[kaindex][kbindex]*d2YCOVijGkb;

349
}
// kbindex = kaindex
kb = index[kbindex];
datatypeb = int(Data.d[nTYPE][kb]);

if (datatypeb == 0)
{
dYCOVipartsGkb = scalar*( dGai

+ dGbi*(-1/(2*pow(((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb)),0.5)))

*((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb))*(1/Gb)

+ dYMeani[kbindex]*(-1/pow(((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb)),0.5))

*(YMean[kbindex]-0.5)/Gb);
dYCOVjpartsGkb = scalar*( dGaj

+ dGbj*(-1/(2*pow(((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb)),0.5)))

*((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb))*(1/Gb)

+ dYMeanj[kbindex]*(-1/pow(((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb)),0.5))

*(YMean[kbindex]-0.5)/Gb);
d2YCOVijGkb = scalar*( d2Gaij

+ d2Gbij *(1/Gb)* ( (-1/(4*pow(((0.25/Gb)-(pow((YMean[kbindex]-


0.5),2)/Gb)),1.5)))*pow(((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb)),2) +
(0.5 * pow(((0.25/Gb)-(pow((YMean[kbindex]-0.5),2)/Gb)),0.5) ) )

+ dGbi*dYMeanj[kbindex] *(1/Gb)* ( (-1/(2*pow(((0.25/Gb)-


(pow((YMean[kbindex]-0.5),2)/Gb)),1.5)))*((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb))*((YMean[kbindex]-0.5)/Gb) + (1/(pow(((0.25/Gb)-
(pow((YMean[kbindex]-0.5),2)/Gb)),0.5)))*((YMean[kbindex]-0.5)/Gb) )

+ dYMeani[kbindex]*dGbj *(1/Gb)* ( (-1/(2*pow(((0.25/Gb)-


(pow((YMean[kbindex]-0.5),2)/Gb)),1.5)))*((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb))*((YMean[kbindex]-0.5)/Gb) + (1/(pow(((0.25/Gb)-
(pow((YMean[kbindex]-0.5),2)/Gb)),0.5)))*((YMean[kbindex]-0.5)/Gb) )

350
+ dYMeani[kbindex]*dYMeanj[kbindex] * ( ((-1/(pow(((0.25/Gb)-
(pow((YMean[kbindex]-0.5),2)/Gb)),1.5)))*( pow((YMean[kbindex]-
0.5),2)/pow(Gb,2) )) - (1/( pow(((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb)),0.5)*Gb)) )

+ d2YMeanij[kbindex] * (-1/pow(((0.25/Gb)-(pow((YMean[kbindex]-
0.5),2)/Gb)),0.5))

*(YMean[kbindex]-0.5)/Gb );
}
else
{
dYCOVipartsGkb = scalar*( dDai

+ dDbi*pow(YMean[kbindex],0.5)

+ dYMeani[kbindex]*0.5*Db/pow(YMean[kbindex],0.5));
dYCOVjpartsGkb = scalar*( dDaj

+ dDbj*pow(YMean[kbindex],0.5)

+ dYMeanj[kbindex]*0.5*Db/pow(YMean[kbindex],0.5));
d2YCOVijGkb = scalar*( d2Daij

+ d2Dbij*pow(YMean[kbindex],0.5)

+ dDbi*dYMeanj[kbindex] * (1/(2*pow(YMean[kbindex],0.5)))

+ dYMeani[kbindex]*dDbj * (1/(2*pow(YMean[kbindex],0.5)))

+ dYMeani[kbindex]*dYMeanj[kbindex]*(-
0.25)*Db/pow(YMean[kbindex],1.5)

+ d2YMeanij[kbindex] * 0.5*Db/pow(YMean[kbindex],0.5) );
}

dYCOVipartsRkakb = 0.0;
dYCOVjpartsRkakb = 0.0;
dYCOVi.a[kaindex][kbindex] =
dYCOVipartsGka*YCOV.R.a[kaindex][kbindex]*YCOV.G.xptr[kbindex]

351
+
YCOV.G.xptr[kaindex]*dYCOVipartsRkakb*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*YCOV.R.a[kaindex][kbindex]*dYCOVipartsGkb;
dYCOVj.a[kaindex][kbindex] =
dYCOVjpartsGka*YCOV.R.a[kaindex][kbindex]*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*dYCOVjpartsRkakb*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*YCOV.R.a[kaindex][kbindex]*dYCOVjpartsGkb;
d2YCOVijRab = 0.0;
d2YCOVij.a[kaindex][kbindex] =
d2YCOVijGka*YCOV.R.a[kaindex][kbindex]*YCOV.G.xptr[kbindex]
+
dYCOVipartsGka*dYCOVjpartsRkakb*YCOV.G.xptr[kbindex]
+
dYCOVipartsGka*YCOV.R.a[kaindex][kbindex]*dYCOVjpartsGkb
+
dYCOVjpartsGka*dYCOVipartsRkakb*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*d2YCOVijRab*YCOV.G.xptr[kbindex]
+
YCOV.G.xptr[kaindex]*dYCOVipartsRkakb*dYCOVjpartsGkb
+
dYCOVjpartsGka*YCOV.R.a[kaindex][kbindex]*dYCOVipartsGkb
+
YCOV.G.xptr[kaindex]*dYCOVjpartsRkakb*dYCOVipartsGkb
+
YCOV.G.xptr[kaindex]*YCOV.R.a[kaindex][kbindex]*d2YCOVijGkb;
}
}

double ModelStructure::CalculateYrhoab(DataStructure &Data, double *x, int ka,


int kb)
{
double rho;
darray tau;
tau.construct(ntheta);

if (ka != kb)
{
tau[0] = fabs(log(Data.d[nstr][ka])-log(Data.d[nstr][kb]));

352
tau[1] = double (Data.d[nTYPE][ka] != Data.d[nTYPE][kb]);
tau[2] = double (Data.d[nspecimen][ka] !=
Data.d[nspecimen][kb]);
tau[3] = double (Data.d[ntest][ka] != Data.d[ntest][kb]);
tau[4] = double (Data.d[npressure][ka] != Data.d[npressure][kb]);

rho = YCOVrho(tau.xptr,x);
}
else
{
rho = 1.0;
}
return(rho);
}

double ModelStructure::YCOVrho(double *tau, double *x)


{
double rho,lnrho,thetanugget;
darray theta;
theta.construct(ntheta);
int i;

thetanugget = exp(x[ithetanugget]);

for (i = 0; i < ntheta; i++)


{
theta[i] = exp(x[itheta[i]]);
}

lnrho = -1.0/thetanugget;

for (i = 0; i < ntheta; i++)


{
lnrho = lnrho + -tau[i]/theta[i];
}

rho = exp(lnrho);
return(rho);
}

double ModelStructure::dYCOVrhoi(int iv, double *tau, double *x)

353
{
double lnrho;
double thetanugget,dthetanugget,dthetanuggeti;
darray theta;
theta.construct(ntheta);
darray dtheta;
dtheta.construct(ntheta);
darray dthetai;
dthetai.construct(ntheta);
double drhoi,dlnrhoi;
int i;

thetanugget = exp(x[ithetanugget]);
dthetanugget = thetanugget;
dthetanuggeti = double(iv == ithetanugget)*dthetanugget;

for (i = 0; i < ntheta; i++)


{
theta[i] = exp(x[itheta[i]]);
dtheta[i] = theta[i];
dthetai[i] = double(iv == itheta[i])*dtheta[i];
}

lnrho = -1.0/thetanugget;
dlnrhoi = 1.0/thetanugget/thetanugget*dthetanuggeti;

for (i = 0; i < ntheta; i++)


{
lnrho = lnrho + -tau[i]/theta[i];

dlnrhoi = dlnrhoi + tau[i]/theta[i]/theta[i]*dthetai[i];


}

drhoi = exp(lnrho)*dlnrhoi;
return(drhoi);
}

double ModelStructure::d2YCOVrhoij(int iv, int jv, double *tau, double *x)


{
double lnrho;
double thetanugget,dthetanugget,dthetanuggeti,dthetanuggetj;

354
double d2thetanugget,d2thetanuggetij;
darray theta;
theta.construct(ntheta);
darray dtheta;
dtheta.construct(ntheta);
darray dthetai;
dthetai.construct(ntheta);
darray dthetaj;
dthetaj.construct(ntheta);
darray d2theta;
d2theta.construct(ntheta);
darray d2thetaij;
d2thetaij.construct(ntheta);
double dlnrhoi,dlnrhoj;
double d2lnrhoij,d2rhoij;
int i;

thetanugget = exp(x[ithetanugget]);
dthetanugget = thetanugget;
dthetanuggeti = double(iv == ithetanugget)*dthetanugget;
dthetanuggetj = double(jv == ithetanugget)*dthetanugget;
d2thetanugget = thetanugget;
d2thetanuggetij = double(iv == ithetanugget)*double(jv ==
ithetanugget)*d2thetanugget;

for (i = 0; i < ntheta; i++)


{
theta[i] = exp(x[itheta[i]]);
dtheta[i] = theta[i];
dthetai[i] = double(iv == itheta[i])*dtheta[i];
dthetaj[i] = double(jv == itheta[i])*dtheta[i];
d2theta[i] = theta[i];
d2thetaij[i] = double(iv == itheta[i])*double(jv ==
itheta[i])*d2theta[i];
}

lnrho = -1.0/thetanugget;;
dlnrhoi = 1.0/thetanugget/thetanugget*dthetanuggeti;
dlnrhoj = 1.0/thetanugget/thetanugget*dthetanuggetj;
d2lnrhoij = -
2*1.0/thetanugget/thetanugget/thetanugget*dthetanuggetj*dthetanuggeti

355
+
1.0/thetanugget/thetanugget*d2thetanuggetij;

for (i = 0; i < ntheta; i++)


{
lnrho = lnrho + -tau[i]/theta[i];
dlnrhoi = dlnrhoi + tau[i]/theta[i]/theta[i]*dthetai[i];
dlnrhoj = dlnrhoj + tau[i]/theta[i]/theta[i]*dthetaj[i];
d2lnrhoij = d2lnrhoij + -
2*tau[i]/theta[i]/theta[i]/theta[i]*dthetaj[i]*dthetai[i]
+ tau[i]/theta[i]/theta[i]*d2thetaij[i];
}

d2rhoij = exp(lnrho)*dlnrhoj*dlnrhoi + exp(lnrho)*d2lnrhoij;


return(d2rhoij);
}

356
REFERENCES
Anderson, D.G. and Woods, R.D. (1975). “Comparison of Field and Laboratory
Moduli,” Proceedings, In Situ Measurement of Soil Properties, ASCE, Vol. 1,
Raleigh, N.C., pp. 66-92.

Anderson, D.G. and Stokoe, K.H., II (1978). “Shear Modulus: A Time Dependent
Property,” Dynamic Geotechnical Testing, ASTM SPT654, ASTM, pp.66-90.

Ang, A.H-S., and Tang, W.H. (1975). Probability Concepts in Engineering


Planning and Design, Volume I - Basic Principles, John Wiley & Sons, New
York, New York, 409 pp.

Ang, A.H-S., and Tang, W.H. (1990). Probability Concepts in Engineering


Planning and Design, Volume II – Decision, Risk and Reliability, John Wiley &
Sons, New York, New York, 562 pp.

Chiara, N. (2001). “Correlation of Small-Strain Dynamic Soil Properties from


Field and Laboratory Studies,” M.S. Thesis, University of Texas at Austin, In
Progress.

Darendeli, M.B. (1997). “Dynamic Properties of Soils Subjected to 1994


Northridge Earthquake,” M.S. Thesis, University of Texas at Austin, 609 pp.

Darendeli, M.B. and Stokoe, K.H., II (1997). “Dynamic Properties of Soils


Subjected to the 1994 Northridge Earthquake,” Geotechnical Engineering Report
GR97-5, Civil Engineering Department, The University of Texas at Austin,
Austin, TX, December.

Darendeli, M.B., Stokoe, K.H., II, Rathje E.M., and Roblee C.J. (2001).
“Importance of Confining Pressure on Nonlinear Soil Behavior and its Impact on
Earthquake Response Predictions of Deep Sites,” XVth International Conference
on Soil Mechanics and Geotechnical Engineering, August 27-31, 2001, Istanbul,
Turkey.

Doroudian, M. and Vucetic, M. (1995). “A Direct Simple Shear Device for


Measuring Small-Strain Behavior,” ASTM Geotechnical Testing Journal, Vol. 18,
No. 1, pp 69-85.

Drnevich, V.P. (1967). “Effect of Strain History on the Dynamic Properties of


Sand,” Ph.D. Dissertation, University of Michigan, 151 pp.

EduPro Civil Systems, Inc. (1998). ProShake Users Manual.

357
Electric Power Research Institute (1993a). “Guidelines for Determining Design
Basis Ground Motions,” Vol. 3; Appendices for Field Investigations, EPRI TR-
102293, Final Report, Palo Alto, CA, November.

Electric Power Research Institute (1993b). “Guidelines for Determining Design


Basis Ground Motions,” Vol. 4; Appendices for Laboratory Investigations, EPRI
TR-102293, Final Report, Palo Alto, CA, November.

Electric Power Research Institute (1993c). “Guidelines for Determining Design


Basis Ground Motions,” Final Report, EPRI TR-102293, Palo Alto, CA,
November.

Fuhriman, Mark, D. (1993). “Crosshole Seismic Tests at Two Northern California


Sites Affected by the 1989 Loma Prieta Earthquake,” M.S. Thesis, The
University of Texas at Austin, 516 pp.

Gilbert, R.B. (1999). “Second-Moment Bayesian Method for Data Analysis in


Decision Making,” Submitted to ASCE Journal of Engineering Mechanics.

Hall, J.R., Jr. and Richart F.E., Jr. (1963).”Effects of Vibration Amplitude on
Wave Velocities in Granular Materials,” Proceedings, Second Pan-American
Conference on Soil Mechanics and Foundations Engineering, Vol. 1, pp. 145-162.

Hardin, B.O. and Music J. (1965). “Apparatus for Vibration of Soil Specimens
During the Triaxial Test,” Symposium on Instrumentation and Apparatus for Soils
and Rocks, ASTM STP 392, ASTM, pp.55-74.

Hardin, B. O. and Drnevich, V.P. (1970a). “Shear Modulus and Damping in


Soils-I: Measurement and Parameter Effects,” Technical Report No. UKY 26-70-
CE2, Soil Mechanics Series No.1, College of Engineering, University of
Kentucky, Lexington, KY, 45 pp.

Hardin, B. O. and Drnevich, V.P. (1970b). “Shear Modulus and Damping in


Soils-II: Design Equations and Curves,” Technical Report No. UKY 27-70-CE3,
Soil Mechanics Series No.2, College of Engineering, University of Kentucky,
Lexington, KY, 49 pp.

Hardin, B. O. and Drnevich, V.P. (1972a). “Shear Modulus and Damping in Soils:
Measurement and Parameter Effects,” Journal of Soil Mechanics and Foundation
Engineering Div., ASCE, Vol. 98 No. SM6, June, pp 603-624.

358
Hardin, B. O. and Drnevich, V.P. (1972b). “Shear Modulus and Damping in
Soils: Design Equations and Curves,” Journal of Soil Mechanics and Foundation
Engineering Div., ASCE, Vol. 98 No. SM7, June, pp 667-692.

Hardin, B. O. (1978). “The Nature of Stress-Strain Behavior of Soils,”


Proceedings, Geotech. Eng. Div. Specialty Conference on Earthquake Eng. and
Soil Dynamics, Vol. 1 ASCE, Pasadena, June, pp. 3-90.

Hwang, S.K. and Stokoe, K.H., II (1993a). “Dynamic Properties of Undisturbed


Soil Samples from Lotung, Taiwan,” Geotechnical Engineering Report GR93-3,
Civil Engineering Department, The University of Texas at Austin, Austin, TX,
March.

Hwang, S.K. and Stokoe, K.H., II (1993b). “Dynamic Properties of Undisturbed


Soil Samples from Treasure Island, California,” Geotechnical Engineering Report
GR93-4, Civil Engineering Department, The University of Texas at Austin,
Austin, TX, March.

Hwang, S.K. and Stokoe, K.H., II (1993c). “Dynamic Properties of Undisturbed


Soil Samples from Gilroy #2, California,” Geotechnical Engineering Report
GR93-5, Civil Engineering Department, The University of Texas at Austin,
Austin, TX, March.

Hwang, S. K. (1997). “Investigation of the Dynamic Properties of Natural Soils,”


Ph.D. Dissertation, University of Texas at Austin, 394 pp.

Idriss, I. M. (1990), “Response of Soft Soil Sites during Earthquakes,”


Proceedings, H. Bolton Seed Memorial Symposium, Vol. 2, May, pp. 273-289.

Isenhower, W.M. (1979). “Torsional Simple Shear / Resonant Column Properties


of San Francisco Bay Mud,” M.S. Thesis, University of Texas at Austin, 307 pp.

Ishibashi, I. and Zhang, X. (1993). “Unified Dynamic Shear Moduli and Damping
Ratios of Sand and Clay,” Soils and Foundations, Vol. 33, No. 1, pp.182-191.

Ishihara, K. (1996). Soil Behavior in Earthquake Geotechnics, Oxford University


Press, Walton Street, Oxford, 350 pp.

Iwasaki, T., Tatsuoka F. and Takagi, Y. (1978). “Shear Moduli of Sands under
Cyclic Torsional Shear Loading,” Soils and Foundations, Vol. 18, No. 1, pp.39-
56.

359
Kokusho, T. (1980). “Cyclic Triaxial Test of Dynamic Soil Properties for Wide
Strain Range,” Soils and Foundations, Vol. 20, No. 2, pp. 45-60.

Kramer, S.L. (1996). Geotechnical Earthquake Engineering, Prentice-Hall, Inc.,


Upper Saddle River, NJ, 653 pp.

Lai, C.G. and Rix, G.J. (1998). “Simultaneous Inversion of Rayleigh Phase
Velocity and Attenuation for Near-Surface Site Characterization,” Georgia
Institute of Technology, School of Civil and Environmental Engineering, Report
No. GIT-CEE/GEO-98-2, July, 258 pp.

Lodde, P.F. (1982). “Shear Moduli and Material Damping of San Francisco Bay
Mud,” M.S. Thesis, University of Texas at Austin.

Masing, G. (1926) “Eigenspannungen und Verfestgung Beim Masing,”


Proceedings, Second International Congress of Applied Mechanics, pp.332-335

Ni, S.-H. (1987). “Dynamic Properties of Sand Under True Triaxial Stress States
from Resonant Column/Torsional Shear Tests,” Ph.D. Dissertation, University of
Texas at Austin, 421 pp.

Richart, J.E., Jr., Hall, J.R., Jr., and Woods, R.O. (1970). Vibrations of Soils and
Foundations, Prentice-Hall Inc., Englewood Cliffs, New Jersey, 414 pp.

Rix, G.J., Lai, C.G., and Spang, A.W., Jr. (2000). “In Situ Measurements of
Damping Ratio Using Surface Waves,” Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, 126, Vol. 5, pp. 472-480.

Schnabel, P.B., Lysmer, J., and Seed, H.B. (1972). “SHAKE: A Computer
Program for Earthquake Response Analysis of Horizontally Layered Sites,”
Report, UCB/EERC-72/12, Univ. of California at Berkeley.

Seed, H.B. and Idriss, I.M. (1970). “Soil Moduli and Damping Factors for
Dynamic Response Analyses,” Report No. EERC-70-10, Earthquake Engineering
Research Center, University of California, Berkeley, CA.

Seed, H.B., Wong, R.T., Idriss, I.M. and Tokimatsu, K. (1986). “Moduli and
Damping Factors for Dynamic Analyses of Cohesionless Soils,” Journal of the
Soil Mechanics and Foundations Division, ASCE, Vol. 112, No. SM11, pp. 1016-
1032.

360
Stokoe, K.H., II and Lodde, P.F. (1978). “Dynamic Response of San Francisco
Bay Mud,” Proceedings, 1980 Offshore Technology Conference, ASCE, Vol. II,
pp. 940-959.

Stokoe, K. H., II, Hwang, S. K., Lee, J. N.-K. and Andrus, R.D. (1994). “Effects
of Various Parameters on the Stiffness and Damping of Soils at Small to Medium
Strains,” Proceedings, International Symposium on Prefailure Deformation
Characteristics of Geomaterials, Vol. 2, Japanese Society of Soil Mechanics and
Foundation Engineering, Sapporo, Japan, September, pp. 785-816.

Stokoe, K.H., II and Darendeli, M.B. (1998). “Laboratory Evaluation of the


Dynamic Properties of Intact Soil Specimens: Garner Valley, California,”
Geotechnical Engineering Report GR98-3, Civil Engineering Department, The
University of Texas at Austin, Austin, TX, March.

Stokoe, K.H., II, Hwang, S.K., Darendeli, M.B. and Lee, N.-K.J. (1998a).
“Correlation Study of Nonlinear Dynamic Soil Properties; Savannah River Site,
Aiken, South Carolina,” Geotechnical Engineering Report GR98-4, Civil
Engineering Department, The University of Texas at Austin, Austin, TX, January.

Stokoe, K.H., II, Moulin, B.S. and Darendeli, M.B. (1998b). “Laboratory
Evaluation of the Dynamic Properties of Intact Soil Specimens from Daniel Island
Terminal, South Carolina,” Geotechnical Engineering Report GR98-2, Civil
Engineering Department, The University of Texas at Austin, Austin, TX,
February.

Stokoe, K.H., II, Moulin, B.S. and Darendeli, M.B. (1998c). “Laboratory
Evaluation of the Dynamic Properties of Intact and Reconstituted Soil Specimens
from the Idaho National Engineering Laboratory,” Geotechnical Engineering
Report GR98-6, Civil Engineering Department, The University of Texas at
Austin, Austin, TX, May.

Stokoe, K.H., II, Darendeli, M.B. and Menq, F.-Y. (1998d). “Laboratory
Evaluation of the Dynamic Properties of Intact Soil Specimens: San-Francisco-
Oakland Bay Bridge, East-Span Seismic Safety Project, Phase I - Site
Characterization,” Geotechnical Engineering Report GR98-8, Civil Engineering
Department, The University of Texas at Austin, Austin, TX, July.

Stokoe, K.H., II, Darendeli, M.B. and Menq, F.-Y. (1998e). “Summary
Laboratory Test Results,” ROSRINE Data Dissemination Workshop, University
of Southern California, Los Angeles, CA, December 15th, 129 pp.

361
Stokoe, K. H., II, Darendeli, M. B., Andrus, R. D. and Brown, L. T. (1999).
“Dynamic soil properties: laboratory, field and correlation studies,” Sêco e Pinto,
(ed.), Proceedings, Second Int. Conf. on Earthquake Geotechnical Engineering,
Lisbon. 21-25 June 1999, (3): pp. 811-845, Rotterdam, Balkema.

Stokoe, K. H., II and Santamarina J.C. (2000). “Seismic-Wave-Based Testing in


Geotechnical Engineering,” Keynote Paper, International Conference on
Geotechnical and Geological Engineering, GeoEng 2000, Melbourne, Australia,
November 19-24, 2000.

Stokoe, K.H., II, Valle, C. and Darendeli, M.B. (2001). “Laboratory Evaluation of
the Dynamic Properties of Coarse Grained Specimens from Various Sites in
California,” Geotechnical Engineering Report GR01-2, Civil Engineering
Department, The University of Texas at Austin, Austin, TX, In Progress.

Sun, J.I., Golesorkhi, R., and Seed, H.B. (1988). “Dynamic Moduli and Damping
Ratios for Cohesive Soils,” Report, UCB/EERC-88/15, Univ. of California at
Berkeley, 48 pp.

Vucetic, M. and Dobry, R. (1991). “Effect of Soil Plasticity on Cyclic Response,”


ASCE, Journal of Geotechnical Engineering, Vol. 117, No. 1, pp. 89-107.

Vucetic, M. (1994). “Cyclic Threshold Shear Strains in Soils,” ASCE, Journal of


Geotechnical Engineering, Vol. 120, No. 12, pp. 2208-2228.

Vucetic, M., Lanzo, G., and Doroudian, M. (1998). “Damping at Small Strains in
Cyclic Simple Shear Test,” ASCE. Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 124, No. 7, pp 585-594.

Zeghal, M., Elgamal A.-W., and Tang, H.T. (1995). “Lotung Downhole Array:
Evaluation of Soil Nonlinear Properties,” Journal of Geotechnical Engineering,
ASCE, 121, Vol. 4, pp. 363-378.

362

Vous aimerez peut-être aussi