Sie sind auf Seite 1von 275

School of Civil and

Environmental Engineering

Simultaneous Inversion
of Rayleigh Phase
Velocity and Attenuation
for Near-Surface Site
Characterization
Carlo G. Lai, PhD
Glenn J. Rix, PhD

National Science Foundation and


U.S. Geological Survey

July 1998

Georgia Institute of Technology


Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology
Georgia Institute of Technology

ACKNOWLEDGMENTS
This research was supported by the National Science Foundation under Grant No.
CMS-9402358 and the U.S. Geological Survey under Award No. 1434-95-G-2634. Any
opinions, findings, and conclusions or recommendations expressed in this material are
those of the authors and do not necessarily reflect the views of the National Science
Foundation and the U.S. Geological Survey. The authors are grateful to Dr. Clifford J.
Astill of the National Science Foundation and Dr. John D. Sims of the U.S. Geological
Survey for their support and encouragement.

ii

TABLE OF CONTENTS
ACKNOWLEDGMENTS

LIST OF TABLES

vii

LIST OF ILLUSTRATIONS

ix

SUMMARY

xvii

CHAPTER
1

INTRODUCTION
1
1.1
Motivation......................................................................................................... 1
1.2
Research Objectives......................................................................................... 4
1.3
Dissertation Outline ........................................................................................ 8

DYNAMIC BEHAVIOR OF SOILS ........................................................................9


2.1
Introduction.......................................................................................................9
2.2
A Survey on Modeling Soil Behavior.......................................................... 10
2.2.1 Overview............................................................................................ 10
2.2.2 The Continuum Mechanics Approach ......................................... 10
2.2.3 The Discrete Mechanics Approach............................................... 13
2.3
Phenomenological Modeling of Soil Behavior.......................................... 15
2.4
Experimental Observations.......................................................................... 16
2.4.1 Overview............................................................................................ 16
2.4.2 Threshold Strains ............................................................................. 16
2.4.3 Stiffness Degradation and Entropy Production.......................... 20
2.5
Constitutive Modeling and Model Parameters.......................................... 26
2.5.1 Overview............................................................................................ 26
2.5.2 Linear Viscoelastic Constitutive Models....................................... 26
2.5.3 Low-Strain Kinematical Properties of Soils (LS-KPS)............... 35
2.5.4 Experimental Measurements of LS-KPS ..................................... 50

RAYLEIGH WAVES IN VERTICALLY HETEROGENEOUS MEDIA ... 57


3.1
Introduction.................................................................................................... 57
3.2
Rayleigh Eigenvalue Problem in Elastic Media ......................................... 58
3.2.1 Solution Techniques ........................................................................ 63
3.3
Effective Rayleigh Phase Velocity in Elastic Media.................................. 67
3.4
Rayleigh Greens Function in Elastic Media.............................................. 71
3.5
Rayleigh Variational Principle in Elastic Media......................................... 76
3.5.1 Modal Rayleigh Phase Velocity Partial Derivatives..................... 78
3.5.2 Effective Rayleigh Phase Velocity Partial Derivatives ................ 83
3.5.3 Attenuation of Rayleigh Waves in Weakly Dissipative
Media.................................................................................................. 87

iii

3.6
3.7
3.8

Rayleigh Eigenvalue Problem in Viscoelastic Media.................................90


3.6.1 A Solution Technique ......................................................................91
Effective Phase Velocity and Greens Function in Viscoelastic
Media ................................................................................................................99
Modal and Effective Partial Derivatives in Viscoelastic Media .............101

SOLUTION OF THE RAYLEIGH INVERSE PROBLEM...........................105


4.1
Introduction ..................................................................................................105
4.2
Ill-Posedness of Inverse Problems ............................................................106
4.3
Coupled Versus Uncoupled Analysis ........................................................107
4.4
Inversion Strategies ......................................................................................108
4.5
Occams Algorithm ......................................................................................110
4.6
Uncoupled Inversion ...................................................................................118
4.6.1 Overview ..........................................................................................118
4.6.2 Uncoupled Fundamental Mode Analysis ....................................120
4.6.3 Uncoupled Equivalent Multi-Mode Analysis..............................121
4.6.4 Uncoupled Effective Multi-Mode Analysis.................................122
4.7
Coupled Inversion........................................................................................123
4.7.1 Overview ..........................................................................................123
4.7.2 Coupled Fundamental Mode Analysis.........................................126
4.7.3 Coupled Equivalent Multi-Mode Analysis ..................................126
4.7.4 Coupled Effective Multi-Mode Analysis .....................................127

RAYLEIGH PHASE VELOCITY AND ATTENUATION


MEASUREMENTS ..................................................................................................129
5.1
Overview........................................................................................................129
5.2
Conventional Measurements Techniques.................................................130
5.2.1 Phase Velocity Measurements.......................................................131
5.2.2 Attenuation Measurements ...........................................................134
5.3
New Measurements Techniques ................................................................138
5.3.1 Uncoupled Measurements .............................................................138
5.3.2 Coupled Measurements..................................................................142
5.4
Statistical Considerations.............................................................................143
5.4.1 Overview ..........................................................................................143
5.4.2 Statistical Aspects of Conventional Measurements...................145
5.4.3 Statistical Aspects of New Measurements Techniques.............147
5.4.3.1 Uncoupled Analysis ........................................................147
5.4.3.2 Coupled Analysis.............................................................148
5.4.4 Statistical Aspects of Uncoupled Rayleigh Inversion ................152
5.4.5 Statistical Aspects of Coupled Rayleigh Inversion.....................154

VALIDATION OF THE ALGORITHMS..........................................................157


6.1
Overview........................................................................................................157
6.2
Lambs Problem............................................................................................157
6.3
Numerical Simulations.................................................................................162

iv

6.3.1
6.3.2
6.3.3

Uncoupled Analyses....................................................................... 171


6.3.1.1 UFUMA Inversion Algorithms.................................... 171
6.3.1.2 UEQMA Inversion Algorithms................................... 178
Coupled Analyses ........................................................................... 188
6.3.2.1 CFUMA Inversion Algorithms .................................... 188
6.3.2.2 CEQMA Inversion Algorithms ................................... 192
Results and Discussion .................................................................. 199

EXPERIMENTAL RESULTS............................................................................... 207


7.1
Overview ....................................................................................................... 207
7.2
Treasure Island Naval Station Site ............................................................ 207
7.3
Uncoupled Inversion................................................................................... 209
7.4
Coupled Inversion ....................................................................................... 216
7.5
Results and Discussion................................................................................ 217

CONCLUSIONS AND RECOMMENDATIONS........................................... 221


8.1
Conclusions................................................................................................... 221
8.2
Recommendations for Future Research................................................... 227

APPENDIX A - Elliptic Hysteretic Loop in Linear Viscoelastic Materials ................. 229


A1 Harmonic Constitutive Relations ......................................................................... 229
A2 Energy Dissipated in Harmonic Excitations ...................................................... 230
A3 Principal Axes of the Elliptic Hysteretic Loop .................................................. 231
APPENDIX B - Effective Rayleigh Phase Velocity Partial Derivatives ........................ 233
APPENDIX C - Description of Computer Codes ........................................................... 241
C1 UFUMA (Uncoupled-Fundamental-Mode-Analysis)....................................... 241
C2 UEQMA (Uncoupled-Equivalent-Multi-Mode-Analysis)................................ 243
C3 CFUMA (Coupled-Fundamental-Mode-Analysis)............................................ 244
C4 CEQMA (Coupled-Equivalent-Multi-Mode-Analysis) .................................... 245
BIBLIOGRAPHY .................................................................................................................. 247

vi

LIST OF TABLES
Number

Page

2.1

Phenomenological Soil Responses to Cyclic Excitation as a Function of


Shear Strain Level ............................................................................................................. 19

2.2

Measurement of Low-Strain Dynamic Properties of Soils (LS-DPS)


Comparison between In-Situ and Laboratory Techniques........................................ 51

6.1

Medium Properties and Frequencies Used for Validation of the Elastic


Lambs Problem.............................................................................................................. 159

6.2

Medium Properties and Frequencies Used for Validation of the


Viscoelastic Lambs Problem........................................................................................ 161

6.3

Medium Properties Used for the Validation of the Inversion Algorithms


(Case 1).............................................................................................................................. 163

6.4

Medium Properties Used for the Validation of the Inversion Algorithms


(Case 2).............................................................................................................................. 163

6.5

Medium Properties Used for the Validation of the Inversion Algorithms


(Case 3).............................................................................................................................. 164

6.6

Inversion Algorithms RMS Error Misfit for Case 1 Soil Profile ............................. 200

6.7

Inversion Algorithms RMS Error Misfit for Case 2 Soil Profile ............................. 202

6.8

Inversion Algorithms RMS Error Misfit for Case 3 Soil Profile ............................. 204

6.9

Inversion Algorithms Performance in Terms of RMS Error Misfit ...................... 204

vii

viii

LIST OF ILLUSTRATIONS
Number

Page

1.1

Seismic Energy Path in Ground Response Analysis (Modified from


EPRI, 1993) ..........................................................................................................................1

1.2

Influence of Gmax on the Acceleration Response Spectrum .........................................2

1.3

Influence of DSmin on the Acceleration Response Spectrum........................................3

2.1

Cause-Effects Relationships in Soil Response to Dynamic Excitations .................. 17

2.2

Dependence of Threshold Shear Strains from Plasticity Index


(After Vucetic, 1994) ........................................................................................................ 20

2.3(a)

Effect of Mean Effective Confining Stress on Modulus Degradation


Curves for Non-Plastic Soils (PI = 0) (After Ishibashi, 1992) .................................. 21

2.3(b)

Effect of Mean Effective Confining Stress on Modulus Degradation


Curves for Plastic Soils (PI = 50) (After Ishibashi, 1992) .......................................... 22

2.4

Modulus Degradation Curves for Soils of Different Plasticity


(After Vucetic and Dobry, 1991) ................................................................................... 23

2.5

Dependence of Energy Dissipated within a Soil Mass on Cyclic Shear


Strain for Soils of Different Plasticity (After Vucetic and Dobry, 1991)................. 23

2.6

Frequency Dependence of the Energy Dissipated Within a Soil Mass


(After Shibuya et al., 1995) .............................................................................................. 25

2.7

Typical Relaxation and Creep Functions for a Viscoelastic Solid............................. 28

2.8

Graphical Representation of the Components of the Complex


Modulus.............................................................................................................................. 39

2.9

Stress-Strain Hysteretic Loop Exhibited by a Linear Viscoelastic


Model during a Harmonic Excitation............................................................................ 40

2.10(a) Influence of Frequency on Phase Velocity of Viscoelastic Waves as


Predicted by the Dispersion Relation Eq. (2.35) ......................................................... 48
2.10(b) Influence of Damping Ratio on Phase Velocity of Viscoelastic Waves
as Predicted by the Dispersion Relation Eq. (2.35)..................................................... 49
2.11

Ranges of Variability of Cyclic Shear Strain Amplitude in Laboratory


and In-Situ Tests (Modified after Ishihara, 1996)........................................................ 52

2.12

Fixed-Free Resonant Column Apparatus (Modified after Ishihara, 1996) .............. 53

3.1

Rayleigh Waves in Vertically Heterogeneous Media ................................................... 61

3.2

Rayleigh Waves Dispersion Curves in Vertically Heterogeneous Media ................. 66


ix

3.3

Rayleigh Displacement Eigenfunctions in Vertically Heterogeneous


Media...................................................................................................................................66

3.4

Geometric Spreading Function for Different Types of Media..................................75

3.5

Partial Derivatives of Rayleigh Phase Velocity with Respect to VP and VS


for an Homogeneous Medium........................................................................................82

3.6

Rayleigh Waves in Viscoelastic Multi-Layered Media..................................................91

3.7(a)

Roots of Rayleigh Secular Function in the Region C of the wR-Plane......................95

3.7(b)

Roots of Rayleigh Secular Function in the Region

4.1

Algorithms for the Solution of the Rayleigh Inverse Problem ................................109

4.2

Flow-Chart of Rayleigh Simultaneous Inversion Using Occams


Algorithm..........................................................................................................................117

4.3

Algorithms for the Solution of the Uncoupled Rayleigh Inverse Problem ...........119

4.4

Algorithms for the Solution of the Strongly Coupled Rayleigh Inverse


Problem.............................................................................................................................125

5.1

Typical Configuration of the Equipment Used in SASW Testing...........................131

5.2

Source-Receivers Configuration in SASW Phase Velocity


Measurements ..................................................................................................................132

5.3(a)

SASW Arrangement Using Common Receiver Midpoint Array .............................133

5.3(b)

SASW Arrangement Using Common Source Array..................................................133

5.4(a)

Attenuation Coefficient Computation at Treasure Island Site.................................136

5.4(b)

Attenuation Coefficient Computation at Treasure Island Site.................................137

5.5(a)

Geometrical Interpretation of Effective Rayleigh Phase Velocity ..........................140

5.5(b)

Geometrical Interpretation of Effective Rayleigh Attenuation


Coefficient ........................................................................................................................141

6.1(a)

Comparison of Solutions for the Elastic Lambs Problem (Case 1)........................159

6.1(b)

Comparison of Solutions for the Elastic Lambs Problem (Case 2)........................160

6.1(c)

Comparison of Solutions for the Elastic Lambs Problem (Case 3)........................160

6.2(a)

Comparison of Solutions for the Viscoelastic Lambs Problem (Case 1)...............161

6.2(b)

Comparison of Solutions for the Viscoelastic Lambs Problem (Case 2)...............161

6.2(c)

Comparison of Solutions for the Viscoelastic Lambs Problem (Case 3)...............162

6.3

Rayleigh Dispersion Curves for Case 1 Soil Profile....................................................164

6.4

Rayleigh Effective Dispersion Curve for Case 1 Soil Profile ....................................165

6.5

Rayleigh Attenuation Curves for Case 1 Soil Profile ..................................................166

of the zR-Plane ....................95

6.6

Rayleigh Dispersion Curves for Case 2 Soil Profile ................................................... 166

6.7

Rayleigh Effective Dispersion Curve for Case 2 Soil Profile.................................... 168

6.8

Rayleigh Attenuation Curves for Case 2 Soil Profile ................................................. 168

6.9

Rayleigh Dispersion Curves for Case 3 Soil Profile ................................................... 169

6.10

Rayleigh Effective Dispersion Curve for Case 3 Soil Profile.................................... 170

6.11

Rayleigh Attenuation Curves for Case 3 Soil Profile ................................................. 170

6.12

Fundamental Mode Theoretical and Synthetic Dispersion Curves


for Case 1 Soil Profile ..................................................................................................... 171

6.13

Shear Wave Velocity Profile from UFUMA Inversion Algorithm


for Case 1 Soil Profile ..................................................................................................... 172

6.14

Convergence of UFUMA Inversion Algorithm for Case 1 Soil


Profile ............................................................................................................................... 172

6.15

Shear Damping Ratio Profile and Theoretical Attenuation Curve


from UFUMA Inversion Algorithm for Case 1 Soil Profile..................................... 173

6.16

Attenuation Curves RMS Misfit Error using UFUMA Inversion


Algorithm for Case 1 Soil Profile.................................................................................. 173

6.17

Fundamental Mode Theoretical and Synthetic Dispersion Curves


for Case 2 Soil Profile ..................................................................................................... 174

6.18

Shear Wave Velocity Profile from UFUMA Inversion Algorithm


for Case 2 Soil Profile ..................................................................................................... 175

6.19

Convergence of UFUMA Inversion Algorithm for Case 2 Soil


Profile ............................................................................................................................... 175

6.20

Shear Damping Ratio Profile and Theoretical Attenuation Curve


from UFUMA Inversion Algorithm for Case 2 Soil Profile..................................... 176

6.21

Attenuation Curves RMS Misfit Error using UFUMA Inversion


Algorithm for Case 2 Soil Profile.................................................................................. 176

6.22

Fundamental Mode Theoretical and Synthetic Dispersion Curves


for Case 3 Soil Profile ..................................................................................................... 177

6.23

Shear Wave Velocity Profile from UFUMA Inversion Algorithm


for Case 3 Soil Profile ..................................................................................................... 177

6.24

Non-Convergence of UFUMA Inversion Algorithm for Case 3 Soil


Profile ............................................................................................................................... 178

6.25

Shear Damping Ratio Profile and Theoretical Attenuation Curve


from UFUMA Inversion Algorithm for Case 3 Soil Profile..................................... 179

6.26

Attenuation Curves RMS Misfit Error using UFUMA Inversion


Algorithm for Case 3 Soil Profile.................................................................................. 179

xi

6.27

Effective Theoretical and Synthetic Dispersion Curves for Case 1


Soil Profile ........................................................................................................................180

6.28

Shear Wave Velocity Profile from UEQMA Inversion Algorithm


for Case 1 Soil Profile......................................................................................................180

6.29

Convergence of UEQMA Inversion Algorithm for Case 1 Soil


Profile................................................................................................................................181

6.30

Shear Damping Ratio Profile and Theoretical Attenuation Curve


from UEQMA Inversion Algorithm for Case 1 Soil Profile ....................................181

6.31

Attenuation Curves RMS Misfit Error using UEQMA Inversion


Algorithm for Case 1 Soil Profile ..................................................................................182

6.32

Effective Theoretical and Synthetic Dispersion Curves for Case 2


Soil Profile ........................................................................................................................183

6.33

Shear Wave Velocity Profile from UEQMA Inversion Algorithm


for Case 2 Soil Profile......................................................................................................183

6.34

Convergence of UEQMA Inversion Algorithm for Case 2 Soil


Profile................................................................................................................................184

6.35

Shear Damping Ratio Profile and Theoretical Attenuation Curve


from UEQMA Inversion Algorithm for Case 2 Soil Profile ....................................184

6.36

Attenuation Curves RMS Misfit Error using UEQMA Inversion


Algorithm for Case 2 Soil Profile ..................................................................................185

6.37

Effective Theoretical and Synthetic Dispersion Curves for Case 3


Soil Profile ........................................................................................................................185

6.38

Shear Wave Velocity Profile from UEQMA Inversion Algorithm


for Case 3 Soil Profile......................................................................................................186

6.39

Convergence of UEQMA Inversion Algorithm for Case 3 Soil


Profile................................................................................................................................186

6.40

Shear Damping Ratio Profile and Theoretical Attenuation Curve


from UEQMA Inversion Algorithm for Case 3 Soil Profile ....................................187

6.41

Attenuation Curves RMS Misfit Error using UEQMA Inversion


Algorithm for Case 3 Soil Profile ..................................................................................187

6.42

Fundamental Mode Theoretical Dispersion and Attenuation


Curves for Case 1 Soil Profile ........................................................................................188

6.43

Shear Wave Velocity and Shear Damping Ratio Profile from


CFUMA Inversion Algorithm for Case 1 Soil Profile................................................189

6.44

Convergence of CFUMA Inversion Algorithm for Case 1 Soil


Profile................................................................................................................................189

xii

6.45

Fundamental Mode Theoretical Dispersion and Attenuation


Curves for Case 2 Soil Profile........................................................................................ 190

6.46

Shear Wave Velocity and Shear Damping Ratio Profile from


CFUMA Inversion Algorithm for Case 2 Soil Profile............................................... 191

6.47

Convergence of CFUMA Inversion Algorithm for Case 2 Soil


Profile ............................................................................................................................... 191

6.48

Fundamental Mode Theoretical Dispersion and Attenuation


Curves for Case 3 Soil Profile........................................................................................ 192

6.49

Shear Wave Velocity and Shear Damping Ratio Profile from


CFUMA Inversion Algorithm for Case 3 Soil Profile............................................... 193

6.50

Convergence of CFUMA Inversion Algorithm for Case 3 Soil


Profile ............................................................................................................................... 193

6.51

Effective Theoretical Dispersion and Attenuation Curves for Case 1


Soil Profile........................................................................................................................ 194

6.52

Shear Wave Velocity and Shear Damping Ratio Profile from


CEQMA Inversion Algorithm for Case 1 Soil Profile .............................................. 195

6.53

RMS Error Misfit of CEQMA Inversion Algorithm for Case 1 Soil


Profile ............................................................................................................ .................. 195

6.54

Effective Theoretical Dispersion and Attenuation Curves for Case 2


Soil Profile........................................................................................................................ 196

6.55

Shear Wave Velocity and Shear Damping Ratio Profile from


CEQMA Inversion Algorithm for Case 2 Soil Profile .............................................. 196

6.56

RMS Error Misfit of CEQMA Inversion Algorithm for Case 2 Soil


Profile ............................................................................................................................... 197

6.57

Effective Theoretical Dispersion and Attenuation Curves for Case 3


Soil Profile........................................................................................................................ 197

6.58

Shear Wave Velocity and Shear Damping Ratio Profile from


CEQMA Inversion Algorithm for Case 3 Soil Profile .............................................. 198

6.59

RMS Error Misfit of CEQMA Inversion Algorithm for Case 3 Soil


Profile ............................................................................................................................... 198

6.60

Inverted Shear Wave Velocity Profiles for Case 1 Soil Stratigraphy ....................... 199

6.61

Inverted Shear Damping Ratio Profiles for Case 1 Soil Stratigraphy...................... 200

6.62

Inverted Shear Wave Velocity Profiles for Case 2 Soil Stratigraphy ....................... 201

6.63

Inverted Shear Damping Ratio Profiles for Case 2 Soil Stratigraphy...................... 201

6.64

Inverted Shear Wave Velocity Profiles for Case 3 Soil Stratigraphy ....................... 203

6.65

Inverted Shear Damping Ratio Profiles for Case 3 Soil Stratigraphy...................... 203

xiii

7.1

Treasure Island National Geotechnical Experimentation Site (After


Spang, 1995) .....................................................................................................................207

7.2

Soil Profile and Properties at the Treasure Island NGES


(After Spang, 1995) .........................................................................................................208

7.3

Fundamental Mode Theoretical and Experimental Dispersion


Curves at Treasure Island NGES .................................................................................210

7.4

Shear Wave Velocity Profile from UFUMA Inversion Algorithm at


Treasure Island NGES ...................................................................................................210

7.5

Convergence of UFUMA Inversion Algorithm at Treasure Island


NGES................................................................................................................................211

7.6

Shear Damping Ratio Profile and Theoretical Attenuation Curve


from UFUMA Inversion Algorithm at Treasure Island NGES ..............................212

7.7

Attenuation Curves RMS Misfit Error using UFUMA Inversion


Algorithm at Treasure Island NGES ...........................................................................212

7.8

Effective Theoretical and Experimental Dispersion Curves at


Treasure Island NGES ...................................................................................................213

7.9

Shear Wave Velocity Profile from UEQMA Inversion Algorithm at


Treasure Island NGES ...................................................................................................213

7.10

Convergence of UEQMA Inversion Algorithm at Treasure Island


NGES................................................................................................................................214

7.11

Shear Damping Ratio Profile and Theoretical Attenuation Curve


from UEQMA Inversion Algorithm at Treasure Island NGES .............................214

7.12

Attenuation Curves RMS Misfit Error using UEQMA Inversion


Algorithm at Treasure Island NGES ...........................................................................215

7.13

Fundamental Mode Theoretical and Experimental Dispersion and


Attenuation Curves at Treasure Island NGES...........................................................215

7.14

Shear Wave Velocity and Shear Damping Ratio Profile from


CFUMA Inversion Algorithm at Treasure Island NGES ........................................216

7.15

Convergence of CFUMA Inversion Algorithm at Treasure Island


NGES................................................................................................................................217

7.16

Comparison at Treasure Island NGES of Shear Wave Velocity


from Surface Wave Test Results with Other Independent
Measurements ..................................................................................................................218

7.17

Comparison at Treasure Island NGES of Shear Damping Ratio


from Surface Wave Test Results with Other Independent
Measurements ..................................................................................................................219

xiv

SUMMARY
Surface wave tests are non-invasive seismic techniques that can be used to determine
the low-strain dynamic properties of a soil deposit. In the conventional interpretation of
these tests, the experimental dispersion and attenuation curves are inverted separately to
determine the shear wave velocity and shear damping ratio profiles at a site. Furthermore,
in the inversion procedure, the experimental dispersion and attenuation curves are matched
with theoretical curves, which include only the fundamental mode of propagation.
The only approach available in the literature that accounts for multi-mode wave
propagation is based on the use of Greens functions where the partial derivatives of
Rayleigh phase velocity with respect to the medium parameters required for the solution of
the inverse problem are computed numerically, and therefore very inefficiently.
This study presents a new approach to the interpretation of surface wave testing. The
new approach is developed around three new ideas. First, the definition of the low-strain
dynamic properties of soils and the Rayleigh wave eigenproblem are revisited and
reformulated within the framework of the linear theory of viscoelasticity. Secondly, an
explicit, analytical expression for the effective Rayleigh phase velocity has been derived.
The effective phase velocity concept forms the basis for the development of a new
surface wave inversion algorithm based on multi-mode rather than modal dispersion and
attenuation curves. Closed-form expressions for the partial derivatives of the effective
Rayleigh phase velocity with respect to the medium parameters have also been obtained by
employing the variational principle of Rayleigh waves.
Thirdly, a numerical technique for the solution of the complex-valued Rayleigh
eigenproblem in viscoelastic media has been implemented. An immediate application of
this solution is the development of a systematic and efficient procedure for simultaneously
determining the shear wave velocity and shear damping ratio profiles of a soil deposit from
the results of surface wave tests. The simultaneous inversion of surface waves data offers
two major advantages over the corresponding uncoupled analysis. First, it explicitly
recognizes the inherent coupling existing between the velocity of propagation of seismic
waves and material damping as a consequence of material dispersion. Secondly, the
simultaneous inversion is a better-posed mathematical problem (in the sense of
Hadamard). The new approach to surface wave analysis is illustrated using several
numerical simulations and experimental data.

xv

xvi

1 INTRODUCTION
1.1 Motivation
Geotechnical earthquake engineering is a well-established discipline concerned with
understanding the role-played by soils in the effects induced by earthquakes. An essential
part of geotechnical earthquake engineering is ground response analysis. The objective of
ground response analysis is the prediction of the free-field site response induced by a
catastrophic event, which may be an earthquake or an explosion, occurring in the interior
of the earths crust. A correct implementation of a ground response analysis requires a
proper modeling of several aspects of the problem including the rupture mechanism at the
source and all the phenomena associated with the propagation of seismic waves from the
source to the desired site at the free-surface. The latter includes transmission of seismic
energy within the continental and oceanic structures of the earth, as well as wave
propagation within the soil mass overlaying the bedrock. Figure 1.1 is a schematic
representation of the spread of seismic energy once it is released from the source. Ground
response analysis has important applications in several areas of geotechnical earthquake
engineering and soil dynamics. Some of the most common include local site response

&&y L ( t )

Local Site Conditions

Regional Geology

Source

&&y F ( t )

Figure 1.1 Seismic Energy Path in Ground Response Analysis (Modified from EPRI,
1993)

Introduction

analyses for the development of design ground motions and response spectra, studies of
soil liquefaction potential, seismic stability analyses of slopes and embankments, and studies
of dynamic soil-structure interaction.
A crucial step in implementing a ground response analysis is the selection of the
constitutive models and their associated parameters used to simulate the dynamic behavior
of the soil. Studies have shown that the strain level(s) induced by an earthquake can range
anywhere from 10-3% up to 1+% depending on several variables including the magnitude of
the event, the source mechanism, the distance from the epicenter, and the properties of
the medium (Kramer, 1996). Therefore, an appropriate constitutive model requires a
definition of the model parameters over a broad range of strain levels, ranging from very
small strain levels (below the linear cyclic threshold strain), where the response of the
medium can be considered linear but not necessarily elastic, to intermediate and large strain
levels where non-linear behavior dominates. In many cases, the very small-strain dynamic
properties of soils are sufficient, since there are often circumstances in seismology and soildynamics where the assumption of linearity is an acceptable approximation.
The following example illustrates the crucial role played by the very small-strain
dynamic properties of a soil deposit in controlling the amplification or de-amplification of
an input motion applied at the bedrock. Figure 1.2 and Figure 1.3 illustrate the results of a
local site response analysis performed using the computer program SHAKE91 (Idriss and
Sun, 1991). This code solves the initial-boundary value problem associated with the one-

1.6

Spectral Acceleration [g]

1.4

Gmax = 150.8 MPa

1.2

Gmax = 67.0 MPa

=5%

1.0

Gmax = 16.8 MPa


Input Motion

0.8
0.6
0.4
0.2
0.0
0.0

0.5

1.0

1.5

2.0

2.5

Period [sec]

3.0

3.5

Figure 1.2 Influence of Gmax on the Acceleration Response Spectrum

4.0

Introduction

dimensional wave propagation in layered viscoelastic media using an equivalent linear


analysis. The input motion used in the numerical simulation was the N90E acceleration
record of the 18 May 1940 El Centro earthquake scaled to a maximum acceleration of
0.15g. This acceleration time history was applied at the base of a homogeneous soil deposit
overlaying the bedrock and having a thickness H = 30 m .
Figure 1.2 shows the influence of the initial tangent shear modulus Gmax (or the shear
wave velocity VS ) on the acceleration response spectrum. As expected, the maximum
response of the spectrum is attained at periods close to the fundamental period of the site,
calculated with the well-known expression 4H / VS .
The influence of the initial shear damping ratio DSmin (value of shear damping ratio
associated with a strain level below the linear cyclic threshold strain) on the acceleration
response spectrum is shown in Figure 1.3. Low values of damping ratio results in a large
amplification of the input motion at the bedrock, particularly at periods close to the natural
period of the site. In both response spectra the structural damping was assumed equal to
5%.

Spectral Acceleration [g]

1.8
1.6

Dsmin = 0.5 %

1.4

Dsmin = 5.5%

1.2

Input Motion

1.0
0.8

=5%

0.6
0.4
0.2
0.0
0.0

0.5

1.0

1.5

2.0

2.5

Period [sec]

3.0

3.5

4.0

Figure 1.3 Influence of DSmin on the Acceleration Response Spectrum


The above figures illustrate the important role played by the low-strain dynamic
properties of a soil deposit in determining the dynamic response of a single degree of
freedom system. The low-strain dynamic properties of soil deposits can be measured with a
variety of techniques. They are generally classified into laboratory techniques and in-situ or

Introduction

field techniques. At the end of Chapter 2, will be presented a summary with the most
important advantages and disadvantages of these techniques.
The main focus of this research effort was on the determination of the very smallstrain dynamic properties of a soil deposit from the interpretation of the results of surface
wave tests. The use of surface (Rayleigh) waves for geotechnical site characterization has
several advantages over more conventional seismic methods such as cross-hole and downhole tests. The most attractive feature of surface wave tests is that they are non-invasive
and hence they do not require the use of boreholes, which permits the tests to be
performed more rapidly and at lower cost than most invasive methods.
Furthermore, at sites where subsurface conditions (e.g. gravelly soils) or environmental
concerns (e.g., solid waste landfills) hinder the use of boreholes and probes, surface wave
tests may constitute the only possible choice for an in-situ site investigation. The next
section describes the most important research objectives pursued during this study.
1.2 Research Objectives
Three primary objectives were envisioned at the beginning of this research effort. The
first objective was the development of a systematic and efficient procedure for
simultaneously determining the low-strain values of VS and DS from the results of surface
wave tests. The most common application of surface wave methods is the determination
of the shear wave velocity profile at a site (Nazarian, 1984; Snchez-Salinero, 1987; Rix,
1988; Stokoe et al., 1989). Recently, Rix et al. (1998a) developed a procedure to calculate
near-surface values of material damping ratio from measurements of the spatial attenuation
of Rayleigh waves. However, until now the two problems of determining the shear wave
velocity and the shear damping ratio profiles at a site have been considered separately and
therefore uncoupled.
One of the goals of this study was to present a different approach to the problem,
where Rayleigh wave phase velocity and attenuation measurements are inverted
simultaneously. The simultaneous inversion of Rayleigh wave phase velocity and
attenuation measurements has two major advantages over the corresponding uncoupled
analysis: it is an elegant procedure to account for the coupling existing between phase
velocity of seismic waves and material damping and the simultaneous inversion is a betterposed mathematical problem (in the sense of Hadamard).
The numerical solution of a non-linear inverse problem is obtained in most cases from
the iterative solution of the corresponding forward problem, which in this case is the
boundary value problem of Rayleigh waves in dissipative media. In developing the solution
of the Rayleigh forward problem, extensive use was made of the powerful and elegant
methods of complex variable theory, more precisely of the theory of analytic functions.

Introduction

Subsequent chapters of this dissertation will provide a description of the theoretical basis
of the simultaneous inversion and will illustrate its applications to some experimental data.
The second objective was, in a sense, motivated by the first objective of this
dissertation. The problem of determining the very small-strain dynamic properties of soils
raises fundamental questions about the meaning of words like properties of soils. Implicit to
the definition of such a term is assumptions of material behavior to which ascribe certain
behavioral properties. As a result, different idealizations of material behavior will require the
definition of different types of material properties. It is unfortunate that often in the
geotechnical literature, it is customary to take for granted certain definitions of material
behavior without ever questioning the validity or the appropriateness of these definitions.
One remarkable example is constituted by the so-called dynamic properties of soils a term used
to collectively denote stiffness and material damping ratio of soils. Chapter 2 of this
dissertation attempts to revisit the definition of these parameters within the framework of
a consistent theory of mechanical behaviour. It is shown that whereas it is not a trivial task
to construct a mathematical model describing the behavior of complex materials such as
soils, it is still possible to formulate relatively simple and accurate phenomenological
models by restricting the formulation to the low strain spectrum. These and other issues
related with constitutive modeling of soils are addressed in this chapter, from a perspective
that is relevant to problems of geotechnical earthquake engineering.
Finally, the third objective of this research effort was developing a better understanding
of the theoretical aspects associated with the interpretation of surface wave measurements.
In the current procedure the shear wave velocity and shear damping ratio profiles are
determined from the application of an inversion algorithm to an experimental dispersion
and attenuation curve. Minimization of the distance (specified by an appropriate definition
of norm) between these curves and those predicted theoretically from an assumed profile
of model parameters is the most common criterion used for the solution of the inverse
problem of surface waves. This procedure has an important limitation: the simulated
(theoretical) dispersion and attenuation curves are defined as modal response functions, i.e. they
are referred to a specific mode of propagation of Rayleigh waves. Conversely, the
experimental dispersion and attenuation curves reflect, in general, the contributions of
several modes of Rayleigh wave propagation and also of body waves in the near field.
There are currently two procedures used to overcome this limitation. The first and
most common one is based on comparing the experimental dispersion and attenuation
curves with those of the fundamental mode obtained theoretically. This method is referred
to in the literature as a 2-D analysis of surface waves (Rosset et al., 1991;). The results
provided by the 2-D analysis are generally satisfactory for normally dispersive (i.e. regular)
shear wave velocity profiles (Gucunski and Woods, 1991; Tokimatsu, 1995). The second
method of interpretation of surface wave data referred in the literature as a 3-D analysis
consists of reproducing with a numerical simulation the actual set-up of the experiment.
The theoretical phase velocities, for instance, are computed from the phase differences

Introduction

between theoretical displacements, and the latter are calculated at locations that emulate
receivers spacings used in the experiment. This method is exact, however it has the
disadvantage of requiring the use of a Greens function program for computing the
displacement field, which is difficult and time-consuming if one wants to include the body
wave contributions in the near field. Furthermore, in this approach the partial derivatives
required for the solution of the non-linear inverse problem of determining an unknown
shear wave velocity profile that corresponds to a given experimental dispersion curve, are
computed numerically. Computation of numerical partial derivatives is notoriously an illconditioned problem, and in this case is also computationally expensive (if compared with
other methods).
This study attempts a new interpretation of surface wave measurements, which
combines the simplicity of a 2-D analysis with the robustness of a 3-D analysis. This is
achieved by deriving an explicit expression for the effective phase velocity of Rayleigh
waves in vertically heterogeneous media (in the literature this quantity is often referred to
as the apparent phase velocity). This is the phase velocity measured experimentally in
surface wave tests if the contribution of the body wave field is neglected. As expected, the
effective phase velocity is a local quantity in the sense that its value varies continuously with
the distance from the source at a given frequency. The effective phase velocity arises from
the superposition of several modes of propagation of Rayleigh waves, each traveling at a
different phase velocity, which is denoted as the modal velocity. In dissipative media, the
effective wave propagation leads naturally to the concept of effective attenuation
coefficient, which is also a local quantity. In light of these results, the commonly used
notions of dispersion and attenuation curves should be more properly replaced by those of
dispersion and attenuation surfaces.
Closed-form analytical expressions for the partial derivatives of the effective phase
velocity with respect to the medium parameters (shear and compression wave velocities)
were also obtained by employing the variational principle of Rayleigh waves. These partial
derivatives are essential for an efficient and accurate solution of the Rayleigh inverse
problem.
Finally, in this attempt to re-formulate the current interpretation of surface wave
measurements, a new approach is proposed which is based on the replacement of the
dispersion and attenuation curves with a different type of response function: the
displacement spectra. The motivation for introducing this new procedure was largely
motivated by the observation that in surface wave tests the primitive quantities measured
experimentally are the displacement phase and amplitudes, and not the Rayleigh phase
velocities and attenuation coefficients. In fact, the effective Rayleigh phase velocity is
nothing but the partial derivative, at constant frequency, of the displacement phase with
respect to the source-receiver distance. A similar interpretation holds for the Rayleigh wave
attenuation coefficient if the notion of displacement phase is replaced by that of
displacement amplitude.

Introduction

In their efforts to identify the structure of the Earth, seismologists use time history
records combined with digital signal processing techniques to obtain modal dispersion and
attenuation curves generated by seismic events. Geotechnical engineers use the dispersive
properties of surface waves generated by active sources for near-surface site
characterization. In attempting to find a solution to their respective problems,
seismologists and geotechnical engineers face similar problems and difficulties, therefore it
is natural that they often come up with similar solution strategies. However, there are two
major differences that profoundly distinguish the problems faced by seismologists and
geotechnical engineers.
The first and most important difference is the scale factor. Whereas for seismologists
the layer thickness of their stratified Earth is on the order of kilometers, geotechnical
engineers deal with layers whose size is two or even three order of magnitude smaller. Also
the frequencies involved in seismology and geotechnical engineering are very different.
Most of the energy contained in a seismic record has a frequency range on the order of 0.1
to 10 Hz. Geotechnical engineers analyze surface waves having frequencies up to 200 Hz
or more. Furthermore, there is a substantial difference in seismology and geotechnical
engineering, concerning the distances over which surface waves are detected and recorded
with seismometers and geophones.
As a result of different spatial and temporal scales involved in seismology and
geotechnical engineering, the phenomenon of surface wave propagation will assume in
these two disciplines certain unique and distinctive features. In seismology for example, the
modes of propagation are in most cases well defined and separated from each other, and
seismologists can determine them from the interpretation of time-history records. On the
contrary, in geotechnical engineering surface wave modes generated by harmonic
oscillators are mostly superimposed rather than separated to each other. It is therefore
natural to expect based on these observations, different methods of interpretation in
seismology and geotechnical engineering.
The second difference between the problems faced by seismologists and geotechnical
engineers is that seismologists do not have control over the source of wave energy:
earthquakes occur at times, locations and with characteristics (duration, frequency content,
source mechanism, etc.) that to this date are not predictable. Conversely, not only can
geotechnical engineers select the source type, but they can also choose its spatial location.
As a result, the task of geotechnical engineers in interpreting surface wave data is
enormously simplified if compared with that of seismologists, as long as the former can
turn to their advantage their ability of control over the source.
In summary, the objectives of this research effort were to reformulate the conventional
interpretation of surface wave tests by developing a technique to simultaneously invert
Rayleigh phase velocity and attenuation data, while accounting for the multi-mode nature

Introduction

of Rayleigh wave propagation in vertically heterogeneous media. These goals were achieved
by first constructing a consistent model of soil dynamic behavior at very-small strain levels.
1.3 Dissertation Outline
The dissertation is organized into eight chapters and three appendices. Chapter 2 is an
introduction to the fundamental problem of modeling soil behavior at low-strain levels
under dynamic excitation. Objective of this chapter is to provide experimental evidence for
supporting the assumption of linear viscoelasticity as an appropriate constitutive law for
modeling dynamic soil behavior at low-strain levels. After a critical overview of the
available models of soil behavior, the viscoelastic constitutive model is introduced and the
corresponding model parameters are rigorously defined. Chapter 3 reviews the theory of
Rayleigh waves propagation in elastic and viscoelastic vertically heterogeneous media. After
illustrating well-known results, an explicit analytical expression for the effective phase
velocity is derived. The variational principle of Rayleigh waves is then used to obtain
explicit relationship for the partial derivatives of the effective Rayleigh phase velocity with
respect to the shear and compression wave velocity of the medium. An important result
presented in this chapter is a new numerical technique for the solution of the complex
Rayleigh eigenproblem in linear viscoelastic media. The technique is quite general and it can
also be applied to strongly dissipative media. Chapter 4 illustrates the main aspects
associated with the solution of the Rayleigh inverse problem, and presents the inversion
algorithms developed in this study. Chapter 5 reviews the conventional techniques used in
surface wave measurements, and introduces a new methodology aimed to improve
consistency, in surface wave testing, between measurement procedures and interpretation
of the results. Some statistical considerations related with surface wave measurements are
also analyzed. Chapter 6 presents the results of a systematic numerical simulation for the
validation of the algorithms developed in this study. Chapter 7 illustrates an example of
application of these algorithms to a real site. Finally, Chapter 8 presents the conclusions of
this research study and illustrates some recommendations for future research.

2 DYNAMIC BEHAVIOR OF SOILS


2.1 Introduction
Scientific understanding proceeds by way of constructing and analyzing models of the segments or
aspects of reality under study. The purpose of these models is not to give a mirror image of reality, not to
include all its elements in their exact sizes and proportions, but rather to single out and make available for
intensive investigation those elements which are decisive. We abstract from non-essentials, we blot out the
unimportant to get an unobstructed view of the important, we magnify in order to improve the range and
accuracy of our observation. A model is, and must be, unrealistic in the sense in which the word is most
commonly used. Nevertheless, and in a sense, paradoxically, if it is a good model it provides the key to
understanding reality. (From Baran and Sweezy, 1968).
Another feature that adds its contribution to the complexity of soil behavior, is the
coupling effect of soil responses. Thermomechanical coupling is one example of a response
interaction effect, which is usually negligible in soils. However, soils may exhibit other
coupling effects, which may be more important including piezo-electric and chemicomechanical coupling (Fam and Santamarina, 1996). Accounting for these response
interaction phenomena may lead to unexpected consequences such as the reformulation of
the principle of effective stress of classical soil mechanics. Newer formulations of this
principle (Mitchell, 1976) recognize that mechanical effects (i.e. change of the effective
stress) may be obtained not only by variations of the gravitational fields (i.e. total stress
and/or hydrostatic pressure), but also by means of electro-chemical perturbations (double
layer theory).
To date, a comprehensive constitutive model able to account in a unified framework for
all these phenomena is not available. Even if such a model existed, its complexity would be
formidable, and most likely not suitable for applications to real-world engineering problems.
However, despite the complexity of soil behavior, soils do not exhibit all their features with
the same degree of importance. Depending on the nature of the problem under
investigation, which includes its intrinsic spatial and temporal scales, the strain levels
involved, and the dominant external fields, many of the features that characterize soil
behavior may be regarded as secondary. They may be interpreted as second or even higher
order effects, and in most cases they may be neglected without appreciable changes in the
resulting analyses.
This is true in many other engineering disciplines and applied sciences, and echoing the
preface of Baran and Sweezy, (1968), it may be said that the art of good engineering often
identifies with the ability of transforming a difficult problem into a simpler one by
attentively discerning what is important from what is superfluous or unessential.
9

10

Dynamic Behavior of Soils

2.2 A Survey on Modeling Soil Behavior


2.2.1

Overview

Two approaches or philosophies are currently used to model the mechanical behavior
of soils and of solids in general. The classical approach is that of continuum mechanics,
which is based on the identification of a deformable medium, in this case soil, with regions
of the three-dimensional Euclidean space. In this approach the mass distribution as well as
all the pertinent field variables (deformation gradient tensor, stress tensor, displacement
vector, etc.) are assumed to be continuous function of the coordinates.
An alternative to continuum mechanics which is gaining popularity is discrete
mechanics, which has its roots in explicitly recognizing the discrete nature of soils (and of
matter in general) and hence modeling them as an aggregate of interacting rigid or
deformable discrete particles.
2.2.2

The Continuum Mechanics Approach

Despite its limitations, continuum mechanics is a formidable tool in the solution of an


innumerable class of practical problems. Most of the strengths of continuum mechanics
come from the consequences of the continuity assumption such as the availability of the
powerful tools of differential and integral calculus (Malvern, 1969). It is by using the
concepts of calculus that the fundamental concepts of stress and strain at a point may be
defined.
Classical continuum mechanics was originally conceived to describe the mechanical
behavior of bodies composed of one constituent which could be solid, liquid or gas as long
as the continuity assumption of the field variables holds (within an acceptable accuracy). But
the assumption of continuity itself does not prevent the possibility of describing the
mechanics of heterogeneous materials. The extension of one-constituent continuum
mechanics to bodies composed of more than a single substance leads naturally to the so
called theories of mixtures (Truesdell, 1957). Although the origin of such theories may be dated
back at the beginning of the century throughout the work of notable chemists and
physicists working on the kinetic theories of gases, the first systematic attempt to construct
a multi-component theory of continuum mechanics is due to Truesdell (1957). Since then
this theory has been extended to include several other features including chemical reactions
occurring among the constituents and electromagnetic effects (Eringen, 1976).
However, classical mixture theories are based on the fundamental postulate that a
mixture is represented by a sequence of continuous bodies all of which occupy the same
regions of space simultaneously (Truesdell, 1957). This assumption of intermiscibility may
be appropriate to model mixtures of fluid-like components; however there are physical
situations where this assumption is not appropriate. A few examples include soils, porous
rocks, granular materials, and multiphase suspensions where the mixture consists of
identifiable solid particles or a matrix surrounded by one or more fluids. These types of

Dynamic Behavior of Soils

11

materials lead to the important distinction between multiphase immiscible mixtures and
miscible mixtures (Goodman and Cowin, 1972). The continuity assumption may still be used
but an additional continuous field variable must be introduced: the volume fraction which
corresponds to the proportion of volume occupied by each component of the mixture.
This scalar field reflects important microstructural features of the mixture subjected to a
thermomechanical process.
Applications of the theory developed by Goodman and Cowin (1972) to model the
behavior of particulate materials have produced interesting results. In their formulation the
balance laws are essentially the same as those proposed by Truesdell (1957) with the
exception that a new equation of balance is included to account for the role-played by
volume fraction changes. This equation is called the balance of equilibrated forces and it
describes the distribution of microstructural forces which is effective in a multiphase
mixture. In essence, the balance of equilibrated forces states that the internal distribution of
forces among the constituents of the mixture is directly related to the changes of their
volume fractions. It can be viewed as a generalization of the principle of effective stress of
classical soil mechanics. One of the most attractive features of this theory is its ability to
model dilatancy, a phenomenon that cannot be modeled with classical continuum
mechanics. Nevertheless, it should be emphasized that although volume fraction is an
important field variable, it is not sufficient to discriminate between two mixtures with
uniform distributions of grains, one with large grains and the other with small grains of the
same material density. In other words, volume fraction alone cannot take into account the
grain size distribution of the constituents (Passman, Nunziato and Walsh, 1984) and in this
sense is a scaleless theory. An interesting new approach to the construction of a multicomponent theory of immiscible mixtures has been proposed recently by Wilmanski (1996).
One of the main features of this theory is the replacement of the equation of balance of
equilibrated forces of Goodman and Cowin with a balance equation for porosity. The
introduction of this new law of balance is motivated by microscopic considerations of the
time rate of change of the geometry of the solid phase of the mixture with respect to the
fluid phase.
The application of the theory of immiscible mixtures to multi-phase media composed of
a solid phase and one or more fluid phases leads naturally to the theories of porous media (Biot,
1955; Bowen, 1982). Such theories, which are special cases of the more general theories
of mixtures, have been the subject of considerable interest over the last 35-40 years. This
interest continues today in the form of different formulations and/or assumptions (De
Boer, 1996). Theories of porous media constitute a possible mathematical framework for
modeling the mechanical behavior of complex multi-phase materials such as soils.
A different line of thought for modeling particulate materials within the realm of
continuum mechanics was developed during the late 1970s and early 1980s by Oda (1978),
Rothenburg (1980), Nemat-Nasser (1982), and Satake (1982), just to mention few of the
early investigators. Their approach was to supplement classical continuum mechanics with

12

Dynamic Behavior of Soils

concepts derived from studies of micromechanics. An important aspect of this theory is


that it provides the link between macroscopic quantities such as the stress tensor and the
microstructural parameters describing the internal arrangement of the particles (such as
particle orientation, orientation of contacts, distribution of inter-particle contact microforces, etc.). This link is obtained by means of appropriate averaging procedures of the
above microstructural variables over a representative elemental volume.
As the studies in this area of micromechanics continued, it became apparent the need
for introducing a new quantity able to describe the spatial arrangements of particulate
materials. This new quantity was introduced with the name of fabric tensor (Oda, 1978), and
since then the use of the fabric tensor as a descriptor of the packing of granular materials
has increased. The fabric tensor is defined as a second rank symmetric tensor and it
describes a continuous field variable. Its use in the mechanics of granular materials has lead
to the important definitions of solid phase and void phase fabric tensors.
These quantities, in particular the void phase fabric tensor, have played a major role in
the applications of the concepts of micromechanics to critical state soil mechanics
(Muhunthan and Chameau, 1996), particularly because it has been shown how to determine
them experimentally (Muhunthan, 1991; Frost and Kuo, 1996; Kuo and Frost, 1997). The
work in this area of soil modeling has been very intense in the recent years, and realistic
constitutive equations relating micro-scale variables and the macro-scale variables have been
proposed for both granular materials (Christoffersen et al., 1981) and cohesive soils (Masad
et al., 1997). The results obtained thus far are encouraging, but more research is required for
a definitive validation of these theories.
This brief survey of the use of a continuum mechanics framework to model soil
behavior is concluded with a short introduction to the so-called polar or generalized theories of
continuum mechanics. It is an interesting subject, which is appealing for its inherent capabilities
of modeling continua having an inner microstructure. The first theory on polar continua
was that of the Cosserat brothers in 1907 who laid down the foundations of what today is
known as Cosserats elasticity to be distinguished from the classical theory of elasticity also
called Cauchys elasticity. Since then, there have been a large number of contributors (Green
and Rivlin, 1964; Eringen and Suhubi, 1964, Mindlin, 1964, Eringen, 1976).
Polar materials are defined as those that admit the existence of couple stresses and body
couples (Truesdell and Noll, 1992). Such a possibility, which is disregarded in classical
continuum mechanics, leads to the construction of an alternative continuum mechanics
where the geometrical points of the continua may possess properties similar to those of
rigid or deformable particles. Thus the geometrical points of classical continuum mechanics
which possess three degrees of freedom are extended to include additional degrees of
freedom which may be the three independent rotations (micropolar continua). In practice this
generalization may continue by simply ascribing additional degrees of freedom to the
material point (Eringen, 1976). Micromorphic continua are defined as media having geometrical

Dynamic Behavior of Soils

13

points with a total of twelve degrees of freedom: three translations, three rotations, and six
microdeformations. Now the material point not only can translate and rotate like a rigid
particle, but it may behave as if it were a deformable particle.
Obviously, the kinematics of polar continua is inherently much more complicated than
the kinematics of classical continuum mechanics, particularly in the case of micromorphic
continua. Non-locality is the peculiar feature of polar continuum mechanics, which essentially
means that this theory is able to account implicitly for the scale effects induced by the inner
microstructure of the continua (Granik and Ferrari, 1993). In this sense classical continuum
mechanics is clearly a scaleless theory.
Polar continuum mechanics is not the only type of non-local continuum theory of
mechanics. Others include the so called materials of grade N (Ferrari et al. 1997) which are
defined as those deformable media whose kinematics are described not only by the
deformation gradient of classical continuum mechanics, but also by higher gradient
measures (Truesdell and Noll, 1992).
Although generalized theories of continuum mechanics have been advanced
considerably in recent years, very few applications have been implemented, particularly in
soil mechanics. Possible explanations include the fact that, despite their elegance, these
theories are complex (Ferrari et al. 1997). Furthermore, there are additional difficulties
associated with the physical interpretation and experimental determination of their
constitutive parameters.
2.2.3

The Discrete Mechanics Approach

The final part of this section is dedicated to alternatives to continuum mechanics as a


framework to model the mechanical behavior of soils. As mentioned at the beginning of
this section, in recent years the popularity of discrete mechanics has increased among soil
mechanicians. The attractive feature of discrete mechanics is the explicit recognition by this
theory of the discrete nature of matter, even though it is clear that the manifestation of this
nature is scale dependent.
By analogy to continuum mechanics, there are several classes of discrete mechanics
theories or techniques. Among them, the two mentioned here are Doublet Mechanics (DM)
and the Discrete (or Distinct) Element Method (DEM). The most popular theory, at least
within the geotechnical community, is certainly the DEM. Its original formulation dates
back to the work of Cundall and Strack in 1979, and since then DEM has been used
extensively in studying the mechanical responses of granular materials (Cundall and Strack,
1979; Ting et al., 1989).
Recently, DEM has also been applied to study the constitutive behavior of watersaturated cohesive soils (Anandarajah, 1996). The essential feature of DEM is modeling a
soil element as a discrete assemblage of interacting rigid or deformable particles. The

14

Dynamic Behavior of Soils

interaction among particles is governed by appropriate constitutive laws, which specify the
magnitude and the direction of the contact forces. The overall system is subjected to the
laws of dynamics with forces and moments due to the self-weight of the particles and to
particle-to-particle interaction.
The computational procedure of DEM involves an explicit time-integration scheme of
the equations of motion. DEM simulations are computationally very expensive and this
limits the size of the problems (number of particles) that can be analyzed. Sometimes for
the purpose of reducing the complexity of the simulations, it has been found convenient to
analyze two-dimensional problems using circular or elliptical disks (Ting et al., 1993). Parallel
computing seems to be the answer for the future of DEM (Kuraoka and Bosscher, 1996),
but more research is needed on modeling the particle-to-particle interactions.
A theory that has been recently proposed and that, in view of the authors (Ferrari et al.,
1997; Granik and Ferrari, 1993), should bridge the gap between continuum and discrete
mechanics (DEM) is Doublet Mechanics (DM). The essential feature of DM is its building
block, which is constituted by a pair of geometric points separated at a finite distance (a
doublet). This elementary unit replaces the differential volume element of continuum
mechanics and the discrete particle or grain of DEM. In the kinematics of DM, the
geometrical points or nodes of a doublet have the following degrees of freedom: they can
move relative to each other in both the axial and the normal directions to their common
axis; moreover, they may rotate about their common axis. DM can be constructed with
different degrees of approximation (Ferrari et al., 1997).
In its simplest form it is a scaleless theory which reduces to classical continuum
mechanics. However, higher order approximations of DM are non-local theories and hence
they may account for the scale effects caused by the discrete nature of the medium.
Preliminary results from the application of this theory are promising. It has been shown for
instance (Granik and Ferrari, 1993), that DM may solve the well-known Flamants paradox
(the Flamants problem is the two dimensional equivalent of the Boussinesqs problem) of
the classical theory of elasticity. However, additional studies and further applications
(supported by experimental data) are required for the ultimate validation of the theory.
This section was not intended to be a comprehensive review of the theories and the
methods currently used to model the mechanical behavior of soils or more in general of
solids. The ones briefly mentioned here are a subset of a much broader class of theories in
continuum and discrete mechanics. However, it is the writers belief that some of the
models presented in this section are of a significative interest in the problem of modeling
soil behavior.

Dynamic Behavior of Soils

15

2.3 Phenomenological Modeling of Soil Behavior


It is apparent from the previous section that the mechanical behavior of soils may be
described with a variety of mathematical idealizations. At the present time none of the
theories that have been cited is able to capture simultaneously all of the features exhibited
by soils, particularly under dynamic excitation and for wide ranges of strain and stress levels.
Each proposed model has its own domain of validity, and it may predict reasonable results
if applied to problems satisfying the assumptions laid at basis of the theory. Often in the
mechanics of materials, the conditions of applicability of a specific theory are dictated by
the intrinsic spatial and temporal scales of a problem. The spatial scale(s) of a problem may
be defined as a measure of the relationships existing among the size of some of its
characteristic elements. Each problem has its own spatial (and temporal) scale which
permits attributing meaning to relative terms such as small and large . The temporal
scale of a problem provides a quantitative description of the relationships existing among
the duration of some of its characteristic temporal events. By specifying the temporal scale
associated to a given problem it becomes possible to attribute a relative meaning to terms
such as fast and slow. All of the problems associated with the mechanical behavior of soils
and other materials are characterized by intrinsic spatial and temporal scales. This statement
can ultimately be justified by the experimental evidence that all natural events are neither
continuous nor instantaneous.
An appropriate assessment of the spatial scales of a problem may show for instance,
that even if there are profound differences between the theories of continuum and discrete
mechanics, the discrepancies between their predictions may be irrelevant for practical
purposes. In seismology, where most of the seismic energy propagates within the frequency
range of about 0.001-100 Hz (Aki and Richards, 1980), the discrete nature of the medium
has no role to play when compared with the lengths of the propagating seismic waves.
Sometimes however, multi-scale phenomena may complicate the analysis of a problem. For
instance in a composite medium characterized by the presence of randomly distributed local
inhomogeneities (scatterers) whose size is comparable with the wavelength of the seismic
waves, a continuous model may be inadequate to represent the scattered wave field.
The time dependent deformation processes exhibited by many rheological materials
including soils are examples of phenomena involving intrinsic temporal scales. For instance,
the process of energy dissipation occurring when a seismic wave propagates within a soil
deposit involves the superposition of several dissipation mechanisms, each characterized by
its own temporal scale (Liu et al., 1976). It is the frequency of excitation that will dictate the
relative importance of the mechanisms activated during the overall process of energy
dissipation (a measure of the amount of unrecoverable energy produced during the
deformation of inelastic materials is the internal entropy density).
Throughout this study the mechanical behavior of soils was modeled using the
phenomenological approach of classical one-constituent continuum mechanics. The
constitutive model used to simulate soil response to dynamic excitations at very small strain

16

Dynamic Behavior of Soils

levels was linear-isotropic viscoelasticity. One of the purposes of the next section is to
justify this choice by illustrating some experimental results.
2.4 Experimental Observations
2.4.1

Overview

During the last 25-30 years, a considerable amount of research has been performed in
an effort to better understand the mechanical response of soils to dynamic excitations.
These studies were performed using a variety of laboratory techniques (e.g., resonant
column tests, cyclic torsional shear tests, cyclic direct simple shear tests, and cyclic triaxial
tests), which allowed researchers to investigate the influence of variables including strain
amplitude and frequency of excitation on soil behavior. The results obtained from this work
have helped in identifying the most important variables and factors affecting the dynamic
behavior of soils.
These variables and factors can be broadly divided into two categories according to their
origin: external variables and intrinsic variables. The external variables correspond to
externally applied actions and include the stress/strain path, stress/strain magnitude,
stress/strain rate, and stress/strain duration depending on the nature of the applied action
(i.e., stress-controlled versus strain-controlled tests). The intrinsic variables correspond to
the inherent characteristics of soil deposits and include the soil type, the size of soil
particles, and the state parameters. The latter include the geostatic effective stress tensor
(which is a measure of the current state of in-situ effective stress), some measure of the
arrangement of soil particles (e.g., the fabric tensor or at least the void ratio, which however
is scale-dependent), and some measure of the stress-strain history (e.g., the yield surface or
at least the preconsolidation pressure). Figure 2.1 summarizes the relationships between
causes and effects in the response of soils to dynamic excitations.
As described in the previous section, soil behavior may be studied using either a
phenomenological or a micromechanical approach. In the phenomenological approach the
main concern is understanding the relationship between causes and effects from a
macroscopic point of view, without attempting an explanation of the observed phenomena
at a microscopic level. This microscopic interpretation is the objective of the
micromechanical approach, which can be implemented by using the framework of either
continuum or discrete mechanics. As already mentioned the approach used in this work to
model the dynamic behavior of soils is phenomenological, and coincides with that of
classical, one-constituent continuum mechanics.
2.4.2

Threshold Strains

Experimental evidence shows that among the external variables affecting soil response
to dynamic excitations, the one that plays the most important role is the magnitude of the
applied stress or strain, or more precisely, the magnitude of the deviatoric strain tensor in

Dynamic Behavior of Soils

Stress/Strain Path
Stress/Strain Magnitude
Stress/Strain Rate
Stress/Strain Duration
Soil Type
Size of Soil Particles
Soil Natural State
- Geostatic Stresses
- Void Ratio/Fabric
- Stress/Strain History

17

External
Causes

Micromechanical

Soil Response
Intrinsic
Causes

Response Functions
Threshold Strains
Stiffness Degradation
Energy Dissipation

Phenomenological

Linear Viscoelastic
Non-Linear Viscoelastic
Non-Linear Elasto-Visco-Plastic

Figure 2.1 Cause-Effect Relationships in Soil Response to Dynamic Excitations


strain controlled tests. This quantity is a measure of the level of shear strains that were
induced in the soil mass during the dynamic excitation. Based on these findings, it was then
possible to define a shear strain spectrum for simple shear conditions where four distinct
types of soil behaviors were identified (EPRI, 1991; Vucetic, 1994).
The very small strain region is defined for values of shear strain in the range 0 < t l
where t l is the so-called linear cyclic threshold shear strain (Vucetic, 1994). Within this region
soil response to cyclic excitation is linear, but not elastic since energy dissipation occurs
even at these very small strain levels (Lo Presti et al., 1997; Kramer, 1996). Although no
stiffness reduction is observed in the soil response for t l (linear behavior), the
hysteretic loop in the stress-strain plane is characterized by a non-null area. The
phenomenon of energy dissipation at very small strain levels is caused by the existence of a
time-lag between say, a driving cyclic strain and the driven cyclic stress in a strain-controlled
test (the word hysteresis comes from the ancient Greek and means lag or delay). This time
lag between cyclic stress and strain is responsible for energy losses over a finite period of
time, which is typical of a viscoelastic behavior. There is little experimental evidence to
support the existence of appreciable phenomena of instantaneous energy dissipation for
t l , which would be typical of an elastoplastic behavior.
Another important feature of soil behavior at very small strain levels is that soil
properties do not degrade as the number of cycles increases, and, as a result, the shape of

18

Dynamic Behavior of Soils

the hysteretic loop does not change with the continued loading (EPRI, 1991; Ishihara,
1996). The value of t l varies considerably with the soil type. For example, t l for sands is
on the order of 10-3%, whereas for normally consolidated clays with a plasticity index (PI) of
50, t l is on the order of 10-2% (Lo Presti, 1987; Lo Presti, 1989).
The small strain region corresponds to shear strain levels in the range t l < t v
where t v is the so-called volumetric threshold shear strain (Vucetic, 1994). The name for this
threshold strain is suggested from the experimental observation that soil response to cyclic
excitation for values of exceeding t v is characterized by irrecoverable changes in
volume in drained tests and development of pore-water pressure in undrained tests
(Vucetic, 1994). This region of the shear strain spectrum is characterized by a non-linear,
inelastic soil response. However, the material properties do not change dramatically with
increasing shear strain, and very little degradation of these properties is observed as the
number of cycles increases (soil hardening or softening). Values of t v , the upper limit for
this region of behavior, are on the order of 5.10-3% for gravels, 10-2% for sands, and 10-1%
for normally consolidated, high plasticity clays (Bellotti et al., 1989; Lo Presti, 1989; Vucetic
and Dobry, 1991).
Values of t v < t pf identify an intermediate strain region where t pf may be called
pre-failure threshold shear strain since values of > t pf characterize soil behavior at large
deformations preceding the conditions at failure. In the intermediate strain region both
instantaneous energy dissipation and energy losses over a finite period of time take place as
the number of cycles progresses. This is mostly due to the irrecoverable microstructural
changes that affect soils once the volumetric cyclic threshold shear strain t v is exceeded
(Vucetic, 1994). In this range of deformation the degradation of soil properties with the
shear strain is apparent not only within the hysteretic loop but also with the increase of the
number of cycles (Ishihara, 1996).
Finally, values of t pf < t f identify the region of large strains (EPRI, 1991; Vucetic,
1994) where soil response to cyclic excitation is highly non-linear and inelastic. This is the
state of soils preceding the condition of failure, which is postulated to occur at the failure
threshold shear strain t f . Table 2.1 shows the shear strain spectrum with the four postulated
types of soil response to cyclic excitation.
Among the four threshold shear strains previously defined, namely t l , t v , t pf , and
t f , two of them are particularly meaningful: the linear cyclic threshold shear strain t l and
the volumetric threshold shear strain t v . The threshold strain t l is important because it
separates the linear (even though inelastic) response from the non-linear response of soils
subjected to cyclic excitations. The volumetric threshold shear strain t v instead, is a

Dynamic Behavior of Soils

19

Table 2.1 Phenomenological Soil Responses to Cyclic Excitation as a Function of Shear


Strain Level

0 < tl

tl < tv

t v < t pf

t pf < t f

Very Small
Strain

Small Strain

Intermediate
Strain

Large Strain

Soil Response

Linear
Viscoelastic

Non-Linear
Viscoelastic

Non-Linear
Elasto-ViscoPlastic

Non-Linear
Elasto-ViscoPlastic

Type of
Non-Linearity

___

Material

Material

Material and
Geometric

Shear Strain
Magnitude

threshold strain used to distinguish between different types of irrecoverable deformations


occurring in soils undergoing harmonic oscillations. For t v it can be postulated that all
the energy losses taking place in a soil specimen are of a viscoelastic nature, i.e. they only
occur over a finite period of time. However, for > t v both phenomena of instantaneous
and finite-time energy dissipation, which is typical of a visco-plastic soil behavior, are
observed experimentally.
The threshold shear strains t l and t v were defined by considering simple shear strain
paths. Soil response to both static and dynamic loadings is strain/stress-path dependent and
hence, different values of t l and t v (and also of t pf , and t f ) would be obtained if
different strain/stress-paths had been used. However the relevance of these concepts and
their implications in understanding the dynamic behavior of soils would be unchanged.
Another factor that affect the values of the threshold shear strains t l and t v is the
mean effective confining pressure which is a measure of the state of effective stresses.
Several studies have shown (Iwasaki et al., 1978; Kokoshu, 1980; Ishibashi and Zhang, 1993;
Ishihara, 1996) that both values of t l and t v increase with increasing confining pressure,
particularly the linear cyclic threshold shear strain t l .
Concerning the influence of the intrinsic properties of soils on the response of these
materials to harmonic excitations, a gradual change has occurred in recent years on how to
approach the problem. Early works (Seed and Idriss, 1970; Hardin and Drnevich, 1972;
Hardin, 1978; Dobry and Vucetic, 1987) treated fine-grained and coarse-grained soils

20

Dynamic Behavior of Soils

separately, and independent correlations were developed for each of these classes. In recent
years (Dobry and Vucetic, 1987; Vucetic and Dobry, 1991; Jamiolkowski et al., 1991;
Vucetic, 1994), it has been recognized that this distinction is unnecessary once variables
such as soil type and size of soil particles are replaced with the plasticity index (PI). The
ability of this index parameter to capture the essential features of soil behavior has been
recognized since the early days of soil mechanics (Casagrande, 1932; Lambe and Whitman,
1969). In the case of the threshold strains, recent research (Vucetic, 1994) has shown that
the plasticity index (PI) plays an important role in determining the magnitude of t l and
t v . Figure 2.2 shows the dependence of t l and t v on the plasticity index.

60

40

Very Small Strains


Small Strains

ge

ge

20

t v

t l

Lin

Lin

30

0.0001

0.001

Av

Av

10
0

era

era

Plasticity Index, PI

50

Intermediate to
Large Strains

0.01

0.1

1.0

Cyclic Shear Strain Amplitude, (%)

Figure 2.2

Dependence of Threshold Shear Strains from Plasticity Index (After


Vucetic, 1994)

The magnitude of both threshold shear strains t l and t v increases with the plasticity
index of the soil. The advantage of using the plasticity index as an independent parameter is
apparent. It provides a unified description of soil properties where the explicit distinction
between fine-grained and coarse-grained becomes unnecessary. It should be remarked
however, that the description of soil properties via the plasticity index is inadequate when
attempting to describe non-plastic soils constituted by large size particles.

Dynamic Behavior of Soils


2.4.3

21

Stiffness Degradation and Entropy Production

So far the emphasis has been placed on the importance of the threshold strains t l and
t v as the primary indicators of the fundamental changes occurring in soil response during
dynamic excitation. From a phenomenological point of view (see Figure 2.1) these changes
result in two observable effects: stiffness reduction and entropy density production (which
is a measure of the amount of energy dissipated within the soil mass during the process of
deformation). The significance of both effects depends strongly on the magnitude of the
shear strains according to the pattern dictated by the threshold strains t l and t v .
Furthermore, both stiffness reduction and entropy density production are also affected by
the mean effective confining pressure and by the soil properties. Figures 2.3(a) and Figure
2.3(b) show the influence of the mean effective confining pressure on stiffness degradation
of soils.

Figure 2.3(a)

Effect of Mean Effective Confining Stress on Modulus Reduction Curves


for Non-Plastic Soils (After Ishibashi, 1992)

It is apparent from these figures that the effect of confining pressure may be significant,
particularly in soils of low plasticity. Entropy density production is also affected by the
effective confining pressure (Ishibashi and Zhang, 1993). Experimental evidence shows that
the energy dissipated in a soil mass decreases as the mean effective confining stress
increases, particularly in low plasticity soils. With regard to the influence of soil type on
stiffness reduction and entropy density production, recent studies (Dobry and Vucetic,

22

Figure 2.3(b)

Dynamic Behavior of Soils

Effect of Mean Effective Confining Stress on Modulus Reduction Curves


for Plastic Soils (After Ishibashi, 1992)

1987; Vucetic and Dobry, 1991) have shown that the plasticity index (PI) is a significant soil
parameter. This is consistent with the strong influence of the PI on the values of the
threshold strains noted above. Figure 2.4 shows the dependence of shear modulus reduction
curves obtained from cyclic laboratory data (Vucetic and Dobry, 1991) on the cyclic shear
strain amplitude and plasticity index of soils. The effects of these parameters on entropy
density production are illustrated in Figure 2.5. Here the amount of energy dissipated within
the soil mass was measured using a different parameter called damping ratio. A rigorous
definition of this quantity will be given later in this chapter.
Other intrinsic soil properties affecting stiffness reduction and entropy density
production include some measure of the initial soil fabric and of the stress/strain history of
the soil deposit. The effect of soil fabric has been quantified only in terms of the initial void
ratio and for very small strain levels (Hardin, 1978; Jamiolkowski et al., 1991). It has been
found that an increase of the void ratio of a soil deposit is accompanied by a decrease of the
stiffness and an increase of the entropy density production at very small strain levels.
The results of laboratory experiments indicate also that the stress/strain history, as
reflected for instance by the yield surface (as a side note, we remark that the overconsolidation
ratio OCR, is a very poor measure of the soil stress history since it corresponds only to one
point of the overall yield surface), affects soil response only at very small strains, where it

Dynamic Behavior of Soils

23

has a moderate influence on soil stiffness of high plasticity soils (Hardin and Drnevich,
1972; Dobry and Vucetic, 1987; Jamiolkowski et al., 1994).

Figure 2.4

Modulus Reduction Curves for Soils of Different Plasticity (After Vucetic


and Dobry, 1991)

Figure 2.5

Dependence of Energy Dissipated within a Soil Mass on Cyclic Shear Strain


for Soils of Different Plasticity (After Vucetic and Dobry, 1991)

24

Dynamic Behavior of Soils

The stress/strain path and the magnitude of the deviatoric strain tensor induced in the
soil mass are the most important external variables, which affect soil response to dynamic
excitations. Other important factors affecting the dynamic behavior of soils include the time
rate of change at which the excitation (stress or strain) is applied and its duration (number
of equivalent cycles). An increasing number of studies (Dobry and Vucetic, 1987; Vucetic
and Dobry, 1991; Shibuya, 1995; Lo Presti et al., 1996; Malagnini, 1996) have investigated
this issue known as the strain-rate effect (in strain-controlled tests). Results from these
studies indicate that the influence of strain-rate effects on soil response is strongly
controlled by the strain level. In particular for clayey soils, the volumetric threshold shear
strain t v increases with increasing strain-rate (Lo Presti et al., 1996).
The stiffness at very small strains ( t l ) does not seem to be affected by the strainrate in low-plasticity soils; however, plastic soils show an increase of the stiffness at very
small strain with increasing strain-rate (Dobry and Vucetic, 1987). At higher strain levels,
experimental evidence shows that the stiffness reduction curves are affected by the strainrate (Dobry and Vucetic, 1987; Lo Presti et al., 1996); soil stiffness generally increases with
increasing strain rate.
Less is known about the influence of the strain-rate effects on the entropy density
production. Even though several studies have attempted to clarify this issue (Dobry and
Vucetic, 1987; Shibuya et al., 1995; Malagnini, 1996; Lo Presti et al., 1997), it is difficult to
draw any general, definitive conclusions. A key parameter controlling entropy density
production is the frequency of excitation. It is observed that at certain frequency bandwiths
the amount of energy dissipated in a soil mass during cyclic excitation is frequency
dependent; experimental evidence also shows the existence of frequency ranges where the
entropy density production is frequency or rate independent (Hardin and Drnevich, 1972;
Shibuya et al., 1995; Lo Presti et al., 1997).
Within this context it is interesting to note that the seismic bandwith, namely the
frequency range (0.001-100 Hz) where most of the seismic energy released during the
earthquakes is concentrated, coincides with the observed zone of frequency independence
of entropy density production (Aki and Richards, 1980; Ben-Menahem and Singh, 1981;
Shibuya et al., 1995). Although it might be expected to observe different frequency
dependence laws at different strain levels, the currently available data do not support this
expectation (Shibuya et al., 1995).
Figure 2.6 shows a conceptual diagram where the amount of energy dissipated
(measured by the material-damping ratio) is plotted as a function of frequency (Shibuya et
al., 1995).
Finally, a factor that may affect soil response to dynamic loads significantly is the
duration of the excitation. In cyclic tests this duration corresponds to the number of
loading cycles. In clayey soils and dry sands, experimental results show that at very small

25

phase A

phase B

phase C

wave loading

seismic loading

traffic loading

max increases

Damping ratio, D

Dynamic Behavior of Soils

0.001

0.01

0.1

10

100

Frequency, f (Hz)

Figure 2.6

Frequency Dependence of the Energy Dissipated within a Soil Mass


(After Shibuya et al., 1995)

strain levels ( t l ), the duration effect on stiffness and entropy density production is
negligible (Shibuya et al., 1995; Lo Presti et al., 1997). As the magnitude of shear strain
increases, material degradation phenomena begin to occur, and the importance of loading
duration becomes more apparent (Lo Presti et al., 1997).
In general it is observed that at large values of shear strains ( > t v ) an increase of the
number of loading cycles yields an acceleration of material deterioration effects such as
stiffness degradation and entropy density production. There are differences, however,
between drained and undrained responses. In sandy soils, stiffness may even increase with
the increase of the number of cycles in drained conditions, whereas the opposite
phenomenon is observed under undrained conditions (Dobry and Vucetic, 1987).
Several other factors may affect stiffness reduction and entropy production during
dynamic excitations. Some of these include cyclic prestraining, creep and relaxation,
anisotropy (structural and stress induced), geological age, diagenetic processes (e.g.
cementation), degree of saturation, and drainage conditions. For most of these factors it is
very difficult to evaluate their effects on stiffness degradation and energy dissipation of soils.
An exception is constituted by anisotropy, where some interesting results can be found in
the literature (Kopperman et al., 1982; Jamiolkowski et al., 1994).

26

Dynamic Behavior of Soils

2.5 Constitutive Modeling and Model Parameters


2.5.1

Overview

The goal of the previous section was to highlight experimental data concerning the
factors (external and intrinsic) affecting the response of soils to dynamic excitation. Shear
strain magnitude is the most important external variable controlling soil response to
dynamic loading. In particular it was shown that for strain-paths of simple shear, it is
possible to define a shear strain spectrum where four different regions of soil response can
be recognized. Table 2.1 illustrates the definition of these four regions and of the associated
threshold shear strains.
Since the primary objective of this research work was the determination of the very
small-strain dynamic properties of soil deposits from the results of surface wave tests, it is
of interest to study the dynamic behavior of soils at strain levels t l . Experimental
evidence shows that in this region of the strain spectrum soils subjected to dynamic
excitations have both the ability to store strain energy (elastic behavior), and to dissipate
strain energy over a finite period of time (viscous behavior). The mechanical behaviour of
such materials can be accurately described from the phenomenological point of view by the
theory of linear viscoelasticity. The following sections illustrate the main features of linear
viscoelastic constitutive models, which include a definition of the general three-dimensional
stress-strain relationships, and of their associated model parameters. The chapter ends with
a section dedicated to the experimental determination of these model parameters.
2.5.2

Linear Viscoelastic Constitutive Models

In purely hypoelastic materials the current state of stress is completely determined by


the current state of strain in a one-to-one or injective correspondence. Viscoelastic materials
have the distinctive feature that the current state of stress is a function not only of the
current state of strain but also of all past states of strains defining the strain-history of the
material under study. There are other types of material responses characterized by this kind
of memory effect (e.g. plastic behaviour), however they are rate independent and hence the
material response to any prescribed strain or stress history does not have any time scale
dependence (Lubliner, 1990).
In the theory of linear viscoelasticity the current state of stress, which is specified by the
stress tensor ij ( t ) , is considered to be related to the past strain history via the following
linear functional (Christensen, 1971) (the summation convention is implied on repeated
indices):

ij ( t ) =

G ijkl ( t )

d kl ( )
d
d

(2.1)

Dynamic Behavior of Soils

27

where kl is the infinitesimal strain tensor and G ijkl is a fourth order tensor-valued function
called the relaxation tensor function of the material. In deriving Eq. 2.1 it was assumed that
the strain history is continuous, however discontinuous strain histories may be represented
as well if the integral appearing in Eq. 2.1 is intended in the Stieltjes sense (Fung, 1965).
Another important assumption required for the derivation of Eq. 2.1 is the time
translation invariance hypothesis, which states that the material response is independent of
any shift along the time axis. The constitutive relationship described by Eq. 2.1 is also called
Boltzmanns equation since it can also be derived by applying the Boltzmanns
superposition principle. The relaxation tensor function G ijkl ( t ) has 81 components;
however, only 21 are independent due to the symmetry of the stress and strain tensors in a
general anisotropic material. Equation 2.1 can be inverted to yield:

ij ( t ) =

J ijkl ( t )

d kl ( )
d
d

(2.2)

where J ijkl ( t ) is a fourth order tensor-valued function called the creep tensor function of
the material. For an isotropic, linear, viscoelastic material the creep and relaxation tensor
functions have only two independent components and they are sufficient to completely
describe the mechanical response of the material. In this case the constitutive relationships
(Eq. 2.1) can be rewritten as:

s ij ( t ) =

kk ( t ) =

2G S ( t )

de ij ( )

3G B ( t )

d kk ( )
d
d

(2.3a)

(2.3b)

1
1
where s ij = ij ij kk and e ij = ij ij kk are the components of the deviatoric
3
3
stress and strain tensors, respectively, and ij is the Kronecker symbol. The scalar functions
G S ( t ) and G B ( t ) are the shear and bulk relaxation functions, respectively. As expected,
shear and volume deformations of viscoelastic isotropic materials are uncoupled, mimicking
a well-known fact of linear isotropic elasticity. From Eq. 2.2 it is possible to obtain
relationships analogous to Eq. 2.3 with the relaxation functions G S ( t ) and G B ( t ) replaced
by the creep functions J S ( t ) and J B ( t ) .
The creep and relaxation functions are material response functions. They are analogous
to the elastic constants in linear elasticity. The important difference is that the creep and

28

Dynamic Behavior of Soils

relaxation functions are no longer constants but time-dependent functions. They both have
an important physical interpretation. The relaxation function G S ( t ) represents the shear
stress response of a material subjected to a shear strain (in a strain-controlled test) specified
as a unit step function (the Heaviside function).

J(t)

G(t)

Time

Time

Figure 2.7 Typical Relaxation and Creep Functions for a Viscoelastic Solid
The creep function J S ( t ) can be viewed as the shear strain response of a material
subjected to a shear stress (in a stress-controlled test) specified as a unit step function.
Analogous interpretations holds for the bulk relaxation and creep functions G B ( t ) and
J B ( t ) . Figure 2.7 illustrates qualitatively the typical shape of a relaxation and a creep
function of a viscoelastic solid.
Often the constitutive relationships of viscoelastic materials are given a physical
interpretation in terms of mechanical models. These are formed by various combinations of
elementary units including linear elastic springs and linear viscous dashpots. The simplest of
these combinations is the Kelvin-Voigt model in which a linear spring and a dashpot are
combined in parallel and the Maxwell model where the above elements are connected in
series. Whereas the former can be used to model certain features of solid behavior, the latter
is suitable to represent fluid behaviour. The relaxation and creep functions of the KelvinVoigt model can be written as follows:

G ( t ) = k H ( t ) + c ( t )

(2.4a)

k
t
H (t)
c
J ( t) =
1 e
k

(2.4b)

Dynamic Behavior of Soils

29

where the index = S, B is a subscript denoting the shear and the bulk mode of
deformation, respectively. The constant k is the stiffness of the elastic spring whereas the
constant c is the coefficient of the viscous dashpot. Finally, H(t) and (t) are the Heaviside
and the Dirac generalized functions. The Kelvin-Voigt model offers a very simplistic and
therefore poor representation of material behaviour. Because J ( 0 + ) = 0 it cannot
represent the instantaneous elastic response that every solid material exhibits when it is
subjected to a suddenly applied stress. Furthermore, although the Kelvin-Voigt model is
able to describe creep, it cannot describe stress relaxation because of the presence of the
Dirac delta function in the expression of the relaxation function G ( t ) .
Better and more accurate mechanical models can be constructed by various
combinations of linear springs and dashpots. The standard linear solid is a relatively simple
model formed by a linear spring and a Kelvin-Voigt unit connected in series. It is able to
describe both phenomena of stress relaxation and instantaneous elastic response in a stresscontrolled test. The relaxation function of the standard linear solid can be written as
follows:

G (t) = G e

+ G g G e e

where G e = G ( t ) =

k 1 k 2
k1 + k 2

(2.5)

, and G g = G ( t = 0 + ) = k 1 are the limiting

values of G ( t ) known as the equilibrium and glassy responses, respectively (Pipkin, 1986);
the terms k 1 and k 2 are the spring constants of the standard linear solid. Finally
=

c
k 1 + k 2

is the relaxation time and it represents the time required by the stress in

the model to reach the equilibrium state represented by G e .


Real materials and in particular soils exhibit more than a single relaxation time, and
therefore the standard linear solid is inadequate to represent their behaviour accurately.
More complicated networks of linear springs and dashpots are required to simulate complex
behaviors characterized by a series of relaxation times. A generalized Maxwell model is
composed of a group of N Maxwell elements in parallel (Malvern, 1969) and the relaxation
function G ( t ) of such a model is given by the following expression:

30

Dynamic Behavior of Soils


N

G ( t) = G e + G i e

t
i

(2.6)

i =1

This equation could be used as a basis for a model fitting procedure (Ferry, 1980) where an
experimental stress relaxation curve ( G ( t ) ) is fitted with the model represented by the
right hand side of Eq. 2.6. The model parameters are the N relaxation times i and the N
amplitudes G i .
Although the use of mechanical models to formulate viscoelastic constitutive models
may be instructive, it is not necessary (Christensen, 1971). Far more general constitutive
relationships may be constructed which are not linked to networks of springs and dashpots.
For example, Eq. 2.6, which represents a model with a discrete spectrum of relaxation
times, may be generalized to a Fredholm integral equation of the first kind representing a
model with a continuous spectrum as follows:
+

G ( t ) = G e + H () e

(2.7)

The function H ( ) is called the relaxation spectrum and it provides important information
about the dissipation mechanisms that may be associated to a spectrum of relaxation times.
The mechanical properties of a viscoelastic material may alternatively be specified by the
relaxation spectrum rather than by the ordinary creep or relaxation functions. However,
solution of the inverse problem represented by Eq. 2.7 may be difficult since the solution of
a Fredholm integral equation of the first kind is a notoriously ill-posed problem (Tikhonov
and Arsenin, 1977).
The complete description of a linear viscoelastic constitutive model requires
specification of two material functions for an isotropic material. These functions may be
either the shear and bulk relaxation functions G S ( t ) and G B ( t ) or the shear and bulk creep
functions J S ( t ) and J B ( t ) . Alternatively, two other response functions may be used: the
shear and bulk relaxation spectra H S ( ) and H B ( ) . Once the two material functions have
been specified, a constitutive relationship such as Eq. 2.3 may be used to compute the
material stress response to any given strain history.
When the prescribed strain or stress history is a harmonic function of time, the
constitutive relationships of the viscoelastic material assume a very simple form. Suppose
for instance that the strain history in Eq. 2.1 is specified by kl ( t ) = 0 kl e it where 0 kl is

Dynamic Behavior of Soils

31

the amplitude of the strain component, i = 1 and is the angular frequency; then the
integral equation (Eq. 2.1) reduces to the following algebraic equation:

ij ( t ) = G *ijkl ( ) 0 kl e it

(2.8)

where G *ijkl ( ) is called the complex tensor modulus and its components are related with
the Fourier sine and cosine transforms of the relaxation tensor function via the following
relations (Christensen, 1971):

G ( 1 ) ijkl ( ) = G ( e ) ijkl + G ijkl ( ) sin d


0

G ( 2 ) ijkl ( ) = G ijkl ( ) cos d


0

(2.9a)

(2.9b)

where G ( 1) ijkl () and G ( 2 ) ijkl () are the real and the imaginary parts of the complex
modulus tensor, respectively, i.e.:

G*ijkl () = G( 1) ijkl () + i G( 2 ) ijkl ()

(2.10)

In Eq. 2.9 it is understood that the relaxation tensor function G ijkl (t ) subtracted from
the equilibrium response G ( e ) ijkl = G ijkl ( t ) is bounded as t in such a way that the
improper integral converges.
From Eq. 2.9 it is clear that the real and the imaginary parts of the complex tensor
modulus are not independent. Their relationship can be easily found to be (Christensen,
1971):

G ( 1) ijkl ( ) = G ( e ) ijkl

2
2 G ( 2 ) ijkl ( )
+
d
0 ( 2 2 )

(2.11)

32

Dynamic Behavior of Soils

This equation is known in the literature (Tschoegl, 1989) as one form of the KramersKrnig relations. It is important because Eq. 2.11 in essence states that viscoelastic materials
are inherently dispersive. Material dispersion is the phenomenon by which the velocity of
propagation of mechanical waves in dissipative media is frequency dependent. It can be
shown (Tschoegl, 1989) that the Kramers-Krnig relation in the form of Eq. 2.11 simply
states that the real and the imaginary part of the complex tensor modulus are the Hilbert
transforms of each other. It can also be proven (Bracewell, 1965) that this result constitutes
the necessary and sufficient condition for the response function G *ijkl () to satisfy the
fundamental principle of causality.
If the material is isotropic the constitutive relationship (2.8) can be rewritten as follows:

s ij ( t ) = 2 G *S ( ) e 0 ij exp( it )

(2.12a)

kk ( t ) = 3G *B ( ) 0 kk exp( i t )

(2.12a)

where e 0 ij and 0 kk are the amplitudes of the deviatoric and volumetric strains, respectively,
and G *S ( ) and G *B () are the complex shear and bulk moduli. One important feature of
the constitutive relationships of linear viscoelastic materials undergoing a steady state
harmonic motion is that stress and strain states are in general out of phase. The amount by
which the stress lags behind the strain is measured by the argument of the complex
modulus, which is also a measure, as shown in the following section, of the amount of
energy dissipated by the material during harmonic excitations. The reciprocal of a complex
modulus is called complex compliance and it is denoted by J * () .
The complex moduli or complex compliances are the fundamental material properties
that need to be specified for the solution of any linear viscoelastic boundary value problem
where all field variables are harmonic functions of time. However, because of the
relationship existing between the complex moduli and the relaxation function (Eq. 2.9 and
its inverse), the complex moduli may also be considered, more generally, as alternative
definitions of mechanical properties of linear viscoelastic materials. Experimental
measurements of the complex moduli may be accomplished by direct observation of the
phase and amplitude relations existing between stress and strain of a material sample
undergoing a cyclic excitation (Ferry, 1980).
In selecting the frequency spectrum to be analyzed, it is important to consider that
information on material behavior over short times may be obtained with high frequency
cyclic experiments, and conversely information on material behaviour over a long time

Dynamic Behavior of Soils

33

period may be obtained by using low frequency tests. However, it should be recognized
that, in principle, knowledge of the relaxation function over a limited time period is not
equivalent to knowing (according to Eq. 2.9) the complex modulus over a finite frequency
range (Christensen, 1971).
In a manner similar to the relaxation function, the experimental determination of a
complex modulus (sometimes called the complex stiffness) G * () requires the selection of
an appropriate constitutive model. This model has to be selected by choosing a satisfactory
compromise between its mathematical complexity and its ability to capture the most
important features of material behaviour. In this sense the description of the mechanical
properties of viscoelastic materials differs profoundly from those of elastic materials. In the
latter case the material properties of an isotropic medium are uniquely specified by two
elastic constants (say the shear and bulk moduli G and B), which may be determined
experimentally from the slopes of appropriate stress-strain curves.
In linear elasticity the difference in response between two materials is simply that one
material is stiffer than the other, either in bulk or in shear or in both modes of deformation.
On the other hand, the mechanical behaviour of a viscoelastic material is more difficult to
characterize. Although the material properties of a linear isotropic viscoelastic material are
uniquely specified by two response functions (say G S ( t ) and G B ( t ) or G *S ( ) and G *B ( ) ),
they may however be constructed in infinitely many ways. In other words, the theory of
linear viscoelasticity allows the construction of a variety of viscoelastic models, with or
without using networks of linear springs and dashpots.
Each of these models has its own features, which may or may not be desirable for
modeling a specific material. Like in any other problem of constitutive modeling, the
selection of a viscoelastic model for fitting the experimental data must be guided by the
material behaviour that the model tries to represent. For example, it has been shown in the
previous section that the amount of energy dissipated by soils during cyclic excitations is
independent from the frequency of excitation, at least within certain frequency bandwidths.
A viscoelastic constitutive model that is intended to simulate soil behaviour under cyclic
loading should be able to reproduce this important feature.
Even though a constitutive model should be as simple as possible, but still able to
reproduce the essential aspects of a material's behavior. A very common viscoelastic model
used to simulate soil behaviour in geotechnical earthquake engineering and soil dynamics is
the Kelvin-Voigt model (Dobry, 1970; EPRI, 1991; Kramer, 1996). Although very simple,
this model fails to reproduce important features of soil or even solid behaviour. It was
shown early in this section that the Kelvin-Voigt model is not able to describe the
instantaneous elastic response displayed by a solid subjected to a suddenly applied stress or
the important phenomenon of stress relaxation.

34

Dynamic Behavior of Soils

Furthermore, the amount of energy dissipated in one cycle of harmonic excitation


predicted by this model is directly proportional to the frequency of vibration ( c );
however, it was pointed out in the previous section that experimental evidence shows that
at very small strain levels energy dissipation in soils is a frequency independent phenomenon
within the seismic band. The current procedure adopted for correcting this inconsistency is
to rearrange the Kelvin-Voigt model to have a viscous dashpot coefficient specified not as a
constant, but as a function inversely proportional to frequency (EPRI, 1991; Kramer, 1996;
Ishihara, 1996). The resulting constitutive model (called a non-viscous or rate-independent
Kelvin-Voigt model by Ishihara (1996)) violates the time-translation invariance hypothesis,
which is one of the postulates of linear viscoelasticity. It can also be shown that when this
rearranged Kelvin-Voigt model is applied to solve the wave equation in linear viscoelastic
materials, it leads to a violation of the elementary principles of causality (Dobry, 1970).
The selection of an oversimplified constitutive model and forcing it to match the
experimental results is rarely a correct strategy to model material behavior. Sometimes the
standard linear solid (with constant dashpot coefficient) is used to replace the Kelvin-Voigt
model for the description of geologic materials (Liu et al., 1976; Jones, 1986). Although a
standard linear solid overcomes the major limitations of the Kelvin-Voigt model, it is still
unable with a single relaxation time to reproduce important features of soil behavior
including the observed frequency independence of energy dissipation within the seismic
band.
A more reasonable approach to the problem of modeling the dynamic behavior of soils
at very small strains would be to assume a material response function (say the relaxation or
the creep function) that is able to capture the essential aspects of the experimental results.
This assumed material response function would in general depend on parameters, which will
be the equivalent of the elastic constants of linear elasticity. These parameters will be
determined by a model fitting procedure applied to some experimental data in a manner
identical to that used when soils are assumed to behave as elastic materials.
This procedure is currently used successfully in modeling other types of viscoelastic
materials such as polymers (Ferry, 1980). It has the great advantage of generality, which
make it suitable to accommodate a large variety of experimental results while remaining
consistent with the fundamental postulates of the theory of linear viscoelasticity.
Examples of application of this approach to geologic materials include the work of Liu
et al. (1976) who assumed a hyperbolic distribution of the relaxation spectrum, and that of
Kjartansson (1979) who adopted a power law time dependence for the creep response
function. Both models are able to predict several features of the behavior of geologic
materials quite accurately including the frequency independence of energy losses in the
seismic band and material dispersion effects.

Dynamic Behavior of Soils


2.5.3

35

Low-Strain Kinematical Properties of Soils (LS-KPS)

In the previous section it was shown that a linear viscoelastic constitutive model for an
isotropic material is completely defined by two response functions which may be given
either in the time or frequency domain. In the time domain they are the relaxation functions
G S ( t ) and G B ( t ) or the creep functions J S ( t ) and J B ( t ) . In the frequency domain they

are the complex moduli G *S ( ) and G *B ( ) or the complex compliances J *S ( ) and J *B () .

Alternatively, the mechanical properties of a viscoelastic material may be also specified


by the relaxation spectra H S ( ) and H B ( ) or by the retardation spectra LS ( ) and L B ( ) ,
which are the equivalent of the complex compliances in the retardation time domain (Ferry,
1980).
As in linear elasticity, knowledge of a pair of response functions in one domain enables
one to determine (at least in principle) the corresponding response functions in any other
domain. Furthermore, as in linear elasticity, the solution of initial-boundary value problems
in linear viscoelasticity requires specification of the material properties only in terms of a
pair of response functions (for dynamic problems, it is also required the knowledge of the
mass density).
Even though the three types of response functions are in a sense equivalent, meaning
that they all contain the same information, it is often advantageous to specify the
mechanical properties of a viscoelastic material by means of the complex moduli. In the
frequency domain, the constitutive relationships of linear viscoelasticity (Eqs. 2.8 and 2.12)
become simple and compact algebraic equations, which resemble those of linear elasticity.
Moreover, in many geotechnical earthquake engineering problems, the dependent variables
are assumed to have an harmonic time dependence or they can be reduced to this case by
using the Fourier transform.
It is a simple matter to show that from the knowledge of the complex moduli one can
easily determine phase velocity and attenuation of harmonic waves propagating in linear
viscoelastic media using the elastic-viscoelastic correspondence principle (Read, 1950; Fung,
1965; Christensen, 1971). According to this principle the solution of a harmonic boundary
value problem in linear viscoelasticity can be easily obtained from the solution of the
corresponding elastic boundary value problem, by extending the validity of the latter
solution to complex values of the field variables. It should be remarked however, that the
validity of the correspondence principle is restricted to problems where the boundaries
conditions (which are the specified stresses and displacements) are time-invariant
(Christensen, 1971).
Application of the elastic-viscoelastic correspondence principle to the equations of
motion (in absence of body forces) governing the propagation of elastic waves in

36

Dynamic Behavior of Soils

unbounded media (Achenbach, 1984), leads to the following pair of complex Helmholtzs
equations:

(div u$ ) =
2

(V )

(curl u$ ) =
2

* 2
P

(V )

* 2
S

div u$

curl u$

(2.13a)

(2.13b)

where 2 denotes the Laplacian operator, the vector u$ = u$ (x, ) is the Fourier
transformed displacement vector ( x is the position vector), and VS* ( ) and VP* ( ) are the
complex S-wave and P-wave velocities, respectively; they govern phase velocity and
attenuation of body waves propagating in linear viscoelastic unbounded media and are
defined by:

V ( ) =

G *B ( ) + 43 G *S

(2.14a)

V ( ) =

G *S ( )

(2.14b)

*
P

*
S

When the equations of motion (Eq. 2.13) are specialized for the two separate cases of
one-dimensional (harmonic) P-wave and S-wave propagation, their general solution can be
written in the following compact form:

u (x, t ) = A e

i * x + t
V

(2.15)

where A is a constant to be determined from the boundary conditions, and = P, S is a


subscript denoting the irrotational (compression) and the equivoluminal (shear) wave

Dynamic Behavior of Soils

37

motion, respectively. In Eq. 2.15 the term V* is the complex wavenumber associated
with the propagation of the -wave. It is convenient to rewrite Eq. 2.15 as follows:

u (x, t ) = Ae

i k x + t

(2.16)

where is the attenuation coefficient and is a measure of the spatial amplitude decay of
the -wave as it propagates through viscoelastic and hence dissipative media. The
term k = V is the (real) wavenumber associated with the -wave propagating with a
(real) phase velocity V . From Eq. 2.14 the attenuation coefficients and the phase velocities
of the irrotational and shear waves are given by the following expressions:

VP ( ) =
*
*
4
G B + 3 G S

P ( ) =
*
*
4
G B + 3 GS

VS ( ) =
*
G S

(2.17a)

(2.17b)

S ( ) =
*
GS

(2.18a)

(2.18b)

In Eqs. 2.17 and 2.18 the symbols ( ) and () denote the real and imaginary parts of a
complex number, respectively.
The mechanics of wave propagation in linear viscoelastic media is completely described
by the phase velocities, VP and VS , and attenuation coefficients, P and S . Whereas VP
and VS give a measure of the speed at which irrotational and equivoluminal disturbances
propagate in a viscoelastic medium, P and S give a description of the spatial attenuation

38

Dynamic Behavior of Soils

of these waves as they propagate through a dissipative material. The velocities of


propagation of P and S waves are directly related to the stiffness of the medium.
On the other hand, the coefficients of attenuation P and S are directly related to the
physical mechanisms responsible for the energy dissipation phenomena occurring within the
material. As shown by the above equations, VP VS , and P , S are material properties

and, as such, they are uniquely determined from the complex moduli G *S ( ) and G *B ( ) ,
and mass density . The other aspect that is important to emphasize is that in a viscoelastic
medium both phase velocities and attenuation coefficients are, in general, frequency
dependent functions. Hence, wave propagation in viscoelastic media gives rise to the
phenomenon of material dispersion as described in Section 2.4.2. One important
consequence of material dispersion is that the shape of the pulse associated with disturbance
changes as it propagates in a viscoelastic medium (Aki and Richards, 1980).
It is instructive to rewrite Eqs. 2.17 and 2.18 in a slightly different form which will make
apparent the close relationship between the coefficients of attenuation and the energy
dissipation phenomena occurring when a wave propagates through a viscoelastic medium.
In particular, by defining the complex constrained modulus G *P ( ) as:

G*P () = (G*B + 43 GS* )

(2.19)

and by separating real and imaginary parts of the complex moduli G *P ( ) and G *S ( ) as:

G * ( ) = G (1) ( ) + i G ( 2) ( )

(2.20)

where again = P, S , it is then possible to rewrite Eqs. 2.17 and 2.18 as follows:

V ( ) =

2 G 2 ( 1) + G 2 ( 2 )

(G

2
( 1)

+G

2
(2)

+ G ( 1)

(2.21a)

Dynamic Behavior of Soils

39

G 2 (1) + G 2 ( 2 ) G ( 1)

( ) =
V ( )
G (2)

(2.21b)

Finally, Eq. 2.21 may be rearranged as:

V ( ) =

G (1)

G 2 (2)
2 1 + 2
G ( 1)

G 2 (2)
1 + 1 + 2
G ( 1)

1+

( ) =

V ( )

G 2 (2)

(2.22a)

G 2 (1)

(2.22b)

G (2)
G ( 1)

The ratio

G (2)
G (1 )

( )

= arg G * = tan ( ) is called the loss tangent or the loss angle, and

G (1) ( ) and G ( 2 ) ( ) are often referred to as the storage and loss moduli, respectively
(Pipkin, 1986). Figure 2.8 shows a graphical representation of these parameters.

( )

G *

G (2)

G *
G

( )

arg G *

G ( 1)

Figure 2.8

( )

G *

Graphical Representation of the Components of the Complex Modulus

40

Dynamic Behavior of Soils


The loss modulus G ( 2 ) ( ) is so named because this parameter is directly related to the

energy dissipated in a viscoelastic material subjected to cyclic loading. It can be shown (see
Appendix A) that the shape of the stress-strain loop predicted by a linear viscoelastic model
during harmonic excitation is elliptical. The equation of the ellipse can be written as follows:
2

G (1)

=1

( 2)
0
0

(2.23)

where the term 0 denotes the amplitude of the harmonic strain . Equation 2.23 is the
equation of an ellipse rotated by an angle ( ) with respect to the strain axis (see Fig. 2.9).
( t)

( ) = 0 eit
( ) = G*
( )

( t)

Figure 2.9

( ) (d ) = G
l

( 2)

Stress-Strain Hysteretic Loop Exhibited by a Linear Viscoelastic Model


during a Harmonic Excitation

An expression of the angle ( ) in terms of the real and imaginary parts of the
complex modulus is given in Appendix A. The elliptical shape of the stress-strain loops
predicted by the theory of linear viscoelasticity compares fairly well with experimental
stress-strain loops obtained for soils at very small strains (i.e. l ) (Dobry, 1970).

Dynamic Behavior of Soils

41

However experimental evidence shows that stress-strain loops obtained at larger strain
amplitudes are cusped (Kjartansson, 1979).
The area enclosed by the ellipse is related to the amount of energy (per unit volume)
dissipated by the material during a cycle of harmonic loading. It can be easily shown (see
Appendix A) that this quantity, Wdissip is equal to:

Wdissip ( ) = G( 2)

(2.24)

and is therefore directly proportional to the loss modulus G ( 2) ( ) . From a


thermodynamic point of view Wdissip is equal to the amount of entropy produced in one
cycle of harmonic loading and due to unrecoverable mechanical work.
At the microscopic level different mechanisms have been proposed (Biot, 1956; Stoll,
1974; Johnston et al., 1979; White, 1983; Leurer, 1997) to explain the process of energy
dissipation occurring at very small strain levels in geological materials subjected to dynamic
excitation. These studies indicate that an interactive combination of several individual
mechanisms is responsible for most of the phenomena macroscopically called energy
dissipation. For coarse-grained soils the two mechanisms that have been postulated to
account for the internal entropy production are frictional losses between soil particles and
fluid flow losses due to the relative movement between the solid and fluid phases. Finegrained soils exhibit more complex phenomena, which are controlled by electromagnetic
interactions between water molecules and microscopic solid particles.
Based on Eq. 2.24, several definitions have been proposed in the literature as measures
of energy dissipation in geological materials (OConnel and Budiansky, 1978; Aki and
Richards, 1980; Ishihara, 1996). Some of them, in particular those used by seismologists,
were inspired by the definitions of energy losses used in other disciplines such as electric
engineering (Cole and Cole, 1941). Most of them are dimensionless parameters proportional
to the ratio between the energy dissipated Wdissip , and some measures of the stored
energy per unit volume.
All these definitions of energy dissipation are consistent with each other only when they
are applied to weakly dissipative viscoelastic materials. In soil dynamics and geotechnical
earthquake engineering, the parameter traditionally used, as a measure of energy dissipation
during harmonic excitation is the so-called material damping ratio:

42

Dynamic Behavior of Soils

D ( ) =
trad

Wdissip ( )

4 Wmax ()

(2.25)

where W max ( ) is the maximum stored energy per unit volume during one cycle of
harmonic excitation.
Although in principle plausible, this definition of material damping ratio is inconvenient
to use. The reason is that the maximum stored energy W max ( ) per unit volume of an
harmonically excited linear viscoelastic material depends not only on the storage modulus
G (1) ( ) , but also on the loss modulus G ( 2) ( ) as well as on their derivatives with respect
to frequency (OConnel and Budiansky, 1978; Tschoegl, 1989).
This interesting result (which is often ignored in the literature of linear viscoelasticity,
see Ferry, 1980; Pipkin, 1986) is caused by the phase lag between the various energy storing
mechanisms which govern the mechanical response of linear viscoelastic materials during
harmonic excitation (Tschoegl, 1989). As a consequence, when the definition of material
damping ratio given by Eq. 2.25 is expressed in terms of the complex modulus G * ( ) , the
ensuing result is very cumbersome. The need for the latter operation is motivated by the
necessity to relate the definition of material-damping ratio to the constitutive parameters of
linear viscoelasticity.
The difficulties associated with the definition of material damping ratio given by Eq.
2.25 can be completely overcome by redefining this parameter, now simply denoted by
D ( ) , as:

D ( ) =

Wdissip ( )

8 Wave ()

(2.26)

where the term Wave ( ) is defined as the average stored energy over one cycle of harmonic
oscillation. An analogous, dimensionless definition of energy dissipation is used by
seismologists, via a material parameter called the quality factor (OConnel and Budiansky,
1978; Aki and Richards, 1980) and denoted by Q ( ) . The two parameters D ( ) and

Q ( ) are related by Q 1 ( ) = 2 D ( ) .

Dynamic Behavior of Soils

43

It can be shown (Tschoegl, 1989) that the term Wave ( ) in the denominator of Eq.
2.26 is equal to W ave ( ) = G ( 1)

4 . Considering this result and that given by Eq.

2.24, Eq. 2.26 yields:

D ( ) =

G( 2)

(2.27)

2 G ( 1)

It is worth noting that when the energy losses in the material are small, the definition of
material damping ratio given by Eq. 2.25 tends to yield the same results as Eqs. 2.26 and
2.27. However, the adoption of the definition given by Eq. 2.26 has the advantage of being
independent from the magnitude of the energy losses. In light of these considerations, the
definition of material damping ratio adopted in this study is that given by Eq. 2.26.
Equation 2.27 can be used to rewrite Eq. 2.22 as follows:

V ( ) =

G (1 )

2 1 + 4 D 2

[1 +

1 + 4 D 2

1 + 4 D 2 1

( ) =

V ( )
2D

(2.28a)

(2.28b)

In geotechnical earthquake engineering, the term low-strain dynamic properties of soils


is used to denote those geotechnical parameters that characterize the dynamic response of
soils at very small strain levels. The most important of these parameters are the stiffness and
the material damping ratio. The symbols commonly used to represent them are G max and
B max for the low-strain shear and bulk stiffnesses, and D Smin and DBmin for the low-strain shear
and bulk material damping ratios. The low-strain dynamic properties are essential
parameters in the solution of many geotechnical earthquake engineering problems. As
described in Chapter 1, they control the response of soil deposits to dynamic excitation, and
hence they play a crucial role in ground response analyses.
However, it is apparent from the above discussion that the low-strain dynamic
properties can be defined more fundamentally in terms of the complex shear and bulk

44

Dynamic Behavior of Soils

moduli G *S ( ) and G *B ( ) , which will be denoted here as the low-strain viscoelastic


properties of soils. Yet, as demonstrated by Equations 2.19, 2.27, and 2.28a, knowledge of
the low-strain dynamic properties of soils is equivalent to specifying the low-strain
viscoelastic properties of soils. In fact these equations provide the means of determining the
real and the imaginary parts of the complex moduli once the material damping ratio and the
stiffness (or equivalently the phase velocity) of the soil have been specified.
These equations can obviously be used also to solve the inverse problem, where the
low-strain dynamic properties of soils are determined from the specified low-strain
viscoelastic properties of soils. Consequently, although the fundamental model parameters
of a linear viscoelastic model in the frequency domain are the low-strain viscoelastic
properties of soils, specification of the low-strain dynamic properties of soils is an
alternative way to provide exactly the same information.
Often in wave propagation problems, it is convenient to replace the low-strain dynamic
properties of soils with the phase velocities VP and VS , and the attenuation coefficients,
P , and S . Equations 2.21 and 2.28 provide the relationships between these quantities,
denoted as the low-strain kinematical properties of soils, and the low-strain viscoelastic
properties of soils and the low-strain dynamic properties of soils.
As already mentioned, the low-strain viscoelastic properties, low-strain dynamic
properties and the low-strain kinematical properties are equivalent. They are simply
alternative ways to characterize the mechanical properties of linear viscoelastic materials.
The choice of which parameters is more appropriate is dictated by the experimental
measurements. In this study, the low-strain kinematical properties are the most frequently
used parameters because in surface wave tests the measured quantities are phase velocities
and attenuation coefficients.
Equation 2.28 give frequency dependent phase velocities and attenuation coefficients,
i.e. the low-strain kinematical properties, for a general linear viscoelastic constitutive model
e
as a function of material damping ratio D ( ) and elastic phase velocity V = G(1) .

These equations can be rewritten in a simpler form as:

V ( ) = Ve 1 D ( )

( ) = k ( ) 2 D ( )

(2.29a)

where k ( ) = / V ( ) and the functions 1 and 2 are defined by:

(2.29b)

Dynamic Behavior of Soils

1 D ( ) =

2 D ( ) =

45

2 1 + 4 D 2

[1 +

1 + 4 D 2

1 + 4 D 2 1
2D

(2.30a)

(2.30b)

At very small strain levels material damping in geological materials is small (Aki and
Richards, 1980; Ishihara, 1996). Therefore it is reasonable to assume:

sup D 01
.

(2.31)

where the symbol sup() denotes the least upper bound of a quantity. Thus it is possible to
expand the functions 1 and 2 of Eq. 2.30 in a Maclaurin series in the variable D . By
retaining only the first and second order terms, Eq. 2.29 simplifies to:

(1 + 2D )

(2.32a)

( ) = k ( ) D ( )

(2.32b)

V ( ) = V

1 + D 2

where the same symbols V ( ) and ( ) have been used for the exact and approximated
expressions of these parameters. Equation 2.32 defines the low-strain kinematical properties
of soils of weakly dissipative (also called low-loss or loss-less) media. The theory of surface
waves in linear viscoelastic media developed in Chapter 3 is referred to weakly dissipative
materials (in this case soil deposits). Thus, in this study the low-strain kinematical properties
of soils are defined by Eq. 2.32.
Application of the correspondence principle to the solution of boundary value
problems in linear viscoelastic media, makes use of a formalism that requires the
introduction of the low-strain kinematical properties of soils as complex-valued phase
velocities V* ( ) . Therefore, as suggested by Eq. 2.16, it is convenient to rewrite Eq. 2.32 in
the following more suitable form:

46

Dynamic Behavior of Soils

V* ( ) =

=
k ( ) + i ( ) k ( ) 1 + iD ( )

(2.33)

Since k ( ) = / V ( ) , Eq. 2.33 can be reduced to:

V* ( ) =

V ( ) 1 iD ( )

[1 + D ( )]
2

(2.34)

For the remainder of this study, Eq. 2.34 will be considered as the formal definition of
low-strain kinematical properties of soils. Equation 2.29a shows that phase velocity V (and
hence stiffness) and material damping ratio D are not independent quantities in linear
viscoelastic media. This fact is also true in weakly dissipative media as shown by Eq. 2.32a.
The functional coupling between soil stiffness and material damping ratio is a direct
consequence of material dispersion, a phenomenon defined in Section 2.5.2 by means of the
Kramers-Krnig relation (Eq. 2.11).
An important corollary of the functional dependence of V upon D is that a
fundamentally correct procedure for the experimental measurement of these parameters
should determine them simultaneously. However, the current practice in geotechnical
engineering is to determine V and D separately. The next section will illustrate the
principles of an experimental procedure to be conducted in laboratory with the resonant
column test where stiffness (or phase velocity) and material damping ratio are determined
simultaneously.
Thus far the derivation of the low-strain kinematical properties has been quite general in
the sense that it has not required any assumption about a specific constitutive model. As a
result the actual formulation, and in particular Eqs. 2.32a and 2.34 are valid for any type of
weakly dissipative, linear viscoelastic solids. (If Eq. 2.31 does not hold, these equations
should be replaced by the first of Eq. 2.33 and Eq. 2.29). It is remarkable to note here that
since Ve is a frequency independent property of a purely elastic material, D ( ) is the only
viscoelastic parameter that needs to be specified for a complete description of low-strain
kinematical properties of soils (i.e. Eq. 2.34).
However, as implied by the Kramers-Krnig relation (Eq. 2.11), the real and imaginary
parts of the complex modulus, G (1) ( ) and G ( 2) ( ) , are not independent, and hence

Dynamic Behavior of Soils

47

from Eq. 2.27 material damping ratio D ( ) cannot be specified arbitrarily. If for example
D ( ) is assumed to be constant and thus frequency independent for the entire frequency

range ] ,+[ , then Eq. 2.32a (and Eq. 2.29a in the general case) would predict that
the velocity of propagation is frequency independent in linear viscoelastic media.
This conclusion is unacceptable not only because it is unsupported by the experimental
data summarized in Section 2.3.3, but also because it violates the fundamental principle of
causality, since no Hilbert transform pair may satisfy the Kramers-Krnig relation with a
constant material damping ratio D (Aki and Richards, 1980). The important result that can
be drawn from this analysis is that even if Eq. 2.34 only requires (apart from Ve ) the

determination of D ( ) , the frequency dependence of D ( ) cannot be prescribed


arbitrarily but it must satisfy the causality constraint imposed by the Kramers-Krnig
relation.
In Section 2.3.3 it was shown that most of the available experimental data indicate that
material damping in soils is a frequency independent phenomenon at very small strain levels
within the seismic frequency band. In the seismological literature (Aki and Richards, 1980;
Kennett, 1983; Keilis-Borok, 1989) it is shown that a nearly constant material damping ratio,
namely a function D ( ) which is frequency independent only over the seismic band,
satisfying the Kramers-Krnig relation yields the following dispersion relation:

V ( ) =

V ( ref )
2 D ref
ln

1 +

(2.35)

where ref denotes an angular reference frequency (usually ref = 2 (1 Hz ) = 2 ).


Equation 2.35 is applicable only for weakly dissipative media within the frequency range
~0.001-10 Hz, (Liu et al., 1976) which corresponds approximately to the seismic band. In
this range of frequencies the material damping ratio D is considered to be a constant and
therefore frequency independent. Frequency or rate independent damping ratio is also
named hysteretic damping ratio (EPRI, 1991; Kramer, 1996); the word hysteretic is often
used in physics and other sciences to denote memory effects processes that are scale
independent (Visintin, 1994).
Theoretically the dispersion relation (Eq. 2.35) could be obtained by imposing the
constraint due to the Kramers-Krnig relation on Eq. 2.32a. This approach, however, turns
out to be very difficult to implement (Tschoegl, 1989). A more practical method (Liu et al.,

48

Dynamic Behavior of Soils

1976) is to assume a creep or a relaxation function which is able to reproduce the nearly
constant D and then use the Kramers-Krnig relation to deduce the dispersion
relationship. An alternative to this approach (Azimi et al., 1968) would be to assume a
Hilbert transform pair which satisfies the Kramers-Krnig relation with a nearly constant
D . Both methods, when applied to weakly dissipative media, yield the dispersion relation
(Eq. 2.35).
It should be apparent from these procedures that postulating the validity of Eq. 2.35 or
any other type of dispersion relation is equivalent to assuming a specific constitutive model
for the material; in fact Eq. 2.35 is an alternative way often used in seismology to specify the
constitutive laws of viscoelastic materials. With the definition of the dispersion relation (Eq.
2.35), the low-strain kinematical properties of soils are completely described by Eq. 2.34
with D ( ) = D = cons tan t .

Normalized Phase Velocity

1.4

1.2

0.8

20

40
60
Frequency [Hz]]

80

100

Damping Ratio = 0.01


Damping Ratio = 0.04
Damping Ratio = 0.08

Figure 2.10(a) Influence of Frequency on Phase Velocity of Viscoelastic Waves as


Predicted by the Dispersion Relation (2.35)
Figure 2.10 illustrates the material dispersion effects predicted by Eq. 2.35. In particular
Fig. 2.10(a) shows the variation of phase velocity with frequency at constant damping ratio

D . In this figure the ratio V ( ) V ref is plotted against frequency with ref = 2 .
As expected, material dispersion effects increase with increasing D . For D = 8 % the

Dynamic Behavior of Soils

49

Normalized Phase Velocity

1.4

1.2

0.8

0.02

0.04
Damping Ratio

0.06

0.08

5 Hz
10 Hz
50 Hz
100 Hz

Figure 2.10(b) Influence of Damping Ratio on Phase Velocity of Viscoelastic Waves as


Predicted by the Dispersion Relation (2.35)
velocity at which a disturbance propagates in a viscoelastic medium may be up to 30%
higher than the corresponding velocity in an elastic medium. Figure 2.10(b) shows an
analogous chart where now phase velocity at constant frequency has been plotted against
damping ratio D .
The description of the low-strain kinematical properties of soils conducted so far has
assumed one-dimensional wave propagation. An important aspect of one-dimensional wave
propagation is that the direction of propagation coincides with the direction of attenuation.
In the case of two or three-dimensional wave propagation, these two directions are not
necessarily the same (Aki and Richards, 1980). By alternatively considering harmonic P and
S-waves, the general solution to Eq. 2.13 (note that for an P-wave curl u = 0 whereas for
an S-wave div u = 0 ) may be written in the form:

u (x , t ) = A e

i k * x + t

(2.36)

50

Dynamic Behavior of Soils

where x is the position vector, A is a constant vector to be determined from the boundary
conditions, and k * = k + i
is the bivector characterizing both the direction of
propagation (specified by the vector k ) and the direction of attenuation (specified by the
vector ) of the -wave. The two vectors k and do not need to be parallel. It is
possible to show (Ben-Menahem and Singh, 1981) that whereas the vector k is normal to
planes of constant phase defined by k x = const. , the vector is normal to planes of
constant amplitude defined by x = const. The phase velocity of the -wave is equal to
V = k . When the vector k is parallel to the vector the corresponding -wave
is called simple (Lockett, 1962).
In a simple -wave the direction of propagation is always the same as the direction of
maximum attenuation. Non simple waves may arise as a result of boundary effects (e.g.
reflection and refraction of harmonic waves at a plane interface) combined with viscoelastic
materials obeying specific constitutive laws (Christensen, 1971). All the propagation
phenomena involving viscoelastic waves considered in this study were assumed to
correspond to simple waves.
2.5.4

Experimental Measurements of Low-Strain Kinematical Properties of Soils

A comprehensive formulation of a model describing material behaviour includes three


fundamental steps. The first step is material modeling which consists of constructing a
mathematical model able to capture the main features of material behaviour (in this case
soil) for a specified range of variation of certain state variables (see preface of Chapter 2). The
second step is constitutive modeling and consists in refining the material modeling with a
specific constitutive law, which is assumed to completely describe the mechanical behaviour
of the material under study. The third step and last step is calibration modeling. This step is
associated with the definition of an appropriate set of experimental procedures for
determining the model parameters of the previously constructed constitutive model.
In the context of this study, the outcome of the first step was the selection of the
phenomenological theory of linear isotropic viscoelasticity to model the dynamic behavior
of soils at very small strain levels. The second step, which was the subject of Sections 2.5.2
and 2.5.3, resulted in the adoption of the dispersion relation (Eq. 2.35) which was largely
based on experimental observations that the material damping ratio for many soils is
hysteretic in nature within the seismic frequency band.
With respect to calibration modeling, it was noted in Section 2.5.3 that specification of
the complex moduli G *S ( ) and G *B ( ) over the entire range of frequencies is sufficient to
completely characterize any isotropic linear viscoelastic constitutive material. The
corresponding relaxation functions G S ( t ) and G B ( t ) in the time domain may be computed
from the inverse Fourier transform of the complex moduli as suggested by Eq. 2.9. A direct

Dynamic Behavior of Soils

51

measurement of the response functions G S ( t ) and G B ( t ) in the time domain is also


possible, although it is not common practice in geotechnical engineering. There are
alternatives to the use of complex moduli as constitutive parameters defining a viscoelastic
model. In soil dynamics and wave propagation problems, the low-strain kinematical
properties often constitute a convenient choice. These are the shear and compression phase
velocities VP and VS , and the shear and compression attenuation coefficients P , and S ,
or equivalently [see Eq. (2.28b) and (2.32b)] the damping ratios D P and D S .
It was emphasized in the previous section that V and D are not independent, and
hence a correct experimental procedure should determine these parameters simultaneously.
Yet, in the current practice of experimental soil dynamics (Kramer, 1996; Ishihara, 1996)
V and D are determined independently from each other. This section will illustrate the
principles of an experimental procedure for simultaneously determine the low-strain dynamic
properties of soils V and D . The technique presented has to be conducted in laboratory,
with a widely known test called the resonant column test. In section 5.3.2 it will be shown how
the simultaneous measurement of the LS-DPS can be performed with the in-situ surface
wave testing.
In the current practice of geotechnical testing, the uncoupled measurement of low-strain
dynamic properties of soils and low-strain kinematical properties of soils can be performed
with a variety of techniques using both in-situ and laboratory tests. In general, laboratory
tests provide more accurate measurements compared with in-situ tests, however they are
not exempt from certain limitations. Table 2.2 attempts to summarize the main differences
between in-situ and laboratory tests.
Table 2.2

Measurement of Low-Strain Dynamic Properties of Soils Comparison


between In-Situ and Laboratory Techniques

Type of Test

Main Advantages

Main Disadvantages

In-Situ Tests

Account for
Scale Factors

Applicable to
any Soil Type

No Alteration
of Soil Natural
State

Difficulties of
Interpretation

Little Control
of Boundary
Conditions

Lack of
General
Standards

Laboratory
Tests

Accurate
Measurements

Repeatable

Controlled
Boundary
Conditions

Very Sensitive
to Sampling
Disturb

Results are
Scale
Dependent

Difficult in
Granular Soils

The field techniques used to measure the low-strain dynamic properties of soils are
mainly of geophysical type because of the ability of these tests to operate at strain levels
10 5 . Some of them like the cross-hole, the down-hole or the seismic cone tests are

52

Dynamic Behavior of Soils

In-Situ Seismic Wave Methods


Resonant Column
Cyclic Triaxial
Cyclic Direct Simple Shear

Earthquakes
10-4

10-3

10-2

10-1

10 0

101

Cyclic Shear Strain, (%)

Figure 2.11

Ranges of Variability of Cyclic Shear Strain Amplitude in Laboratory and


In-Situ Tests (Modified from Ishihara, 1996)

invasive, and hence they require the use of boreholes and probes. Some others like the
seismic reflection, the seismic refraction or the surface wave tests are non-invasive, and they
can be executed from the ground surface without the need of using boreholes or probes.
Concerning the use of laboratory techniques to measure the low-strain dynamic properties
of soils, very few of them can operate at very small strain levels; among them the most
popular is certainly the resonant column which allow the tests to be performed at strain
level < 10 6 .
Figure 2.11 illustrates the strain levels mobilized by the most common in-situ and
laboratory tests. Current techniques used to measure the low-strain dynamic properties of
soils consider soil stiffness and material damping ratio as independent parameters. As a result
each of these parameters is measured separately. Using the results presented in section 2.5.2
and Section 2.5.3 it is possible to show how a laboratory technique such as the resonant
column test can effectively be used to simultaneously determine stiffness and damping ratio of
a soil specimen. Since these parameters are determined at specific frequencies of excitation,
the proposed method is well suited to investigate the frequency dependence laws of these
important soil properties.
In the resonant column test, a solid or hollow circular cylindrical soil specimen is
subjected to harmonic excitation by an electromagnetic driving system (Drnevich, 1985).
The soil specimen can be excited in either the torsional or the longitudinal modes of vibration.

Dynamic Behavior of Soils

53

The study presented hereinafter refers to a stress-controlled loading resonant column test in
the torsional mode of oscillation. The frequency and amplitude of the harmonic excitation
is controlled by the electromagnetic driving system.
Figure 2.12 shows schematically a fixed-free resonant column apparatus, which is fixed
at the base and free to rotate at the top where a driving torque T0 e it is applied. In the
proposed setting of the test, the parameter measured experimentally is the shear complex
modulus G *S ( ) .

T0 e it

Figure 2.12

Fixed-Free Resonant Column Apparatus (Modified after Ishihara, 1996)

Introducing a system of cylindrical coordinates


governing the vibrations of an elastic cylinder is:

2u =

1 2u

VS2 t 2

{r , , z}

the equation of motion

(2.37)

54

Dynamic Behavior of Soils

2
where =

2 1
1 2
2
+
+
+
denotes the Laplacian operator in cylindrical
r 2 r r r 2 2 z 2

coordinates, VS =

GS
is the elastic shear wave velocity, G S is the elastic shear modulus,

and is the mass density of the soil specimen; finally u (r , z , t) is the displacement
component in the direction .
If at z = 0 the cylinder is subjected to a specified harmonic torque T0 e it , the solution
to Eq. 2.37 may be sought in the form:

u (r , z , t ) = (r ) (z ) e it

(2.38)

where is the angular frequency of oscillation. If Eq. 2.38 is substituted into Eq. 2.37, this
second order partial differential equation becomes two ordinary differential equations in the
unknown functions ( r ) and ( z) which can be easily solved. The result is:

z
z
u (r , z , t ) = A sin
r e it
+ B cos

h
h

(2.39)

h 2 2
where 2 =
and h is the height of the cylinder. A and B are two constants to be
GS
determined from the boundary conditions which are:

u (r , h , t ) = 0

z (r ,0 , t )ds = T0 e it

(2.40)

where S = R 2 is the cross sectional area of the cylinder having a radius R.


u
, the twisting moment
z
M(z , t ) is computed by integrating z (r , z , t ) over the cross sectional area to yield:
Since the only non-zero stress component is z = G S

Dynamic Behavior of Soils

M (z , t ) =

R 4 2 h
z
z i t

A cos
B sin
e
h
h
2

55

(2.41)

By applying the boundary conditions in Eq. 2.40, it is found that the ratio between the
u
twisting moment M(z , t ) and the angle of twist = ( z) e it at z = 0 is given by
r
(Christensen, 1971):

M (0 , t )
T0
R 4 2 h
=
=

cot
( 0 ) e i t ( 0 )
2

(2.42)

Application of the elastic-viscoelastic correspondence principle allows one to obtain the


corresponding viscoelastic solution from the elastic solution. The result is:

T0
R 4 2 h
=

cot *
*
(
)
0
2

(2.43)

2 h 2
and G *S ( ) is the complex shear modulus. An inspection of Eq.
G S* ( )
2.43 suggests that experimental measurement of the angle of twist at the top of the
specimen ( 0) will allow the complex modulus G *S ( ) to be determined once the amplitude
of the applied torque T0 and the geometry of the specimen are known.

where * ( ) =

It should be remarked that since the twisting moment M(0, t ) and the angle of twist

( 0) e it will be in general out of phase, ( 0) is a complex number. If this analysis is carried

out over a wide range of frequencies, knowledge of G *S ( ) will permit the determination of
other types of response functions such as the creep and relaxation functions. Finally, the
low-strain dynamic properties of soils in shear are determined from the knowledge of
G *S ( ) using Eq. 2.18a for shear wave velocity, VS ( ) , and Eq. 2.27 for material damping
ratio, D S ( ) .
This analysis can be generalized to other modes of excitation. If for example, the soil
specimen in the resonant column test were excited in the longitudinal direction, the

56

Dynamic Behavior of Soils

procedure just presented would lead to the experimental measurement of the complex
Youngs modulus G *E ( ) .

3 RAYLEIGH WAVES IN VERTICALLY HETEROGENEOUS


MEDIA
3.1 Introduction
The purpose of this chapter is to illustrate the most salient aspects of the theory of
Rayleigh wave propagation in vertically heterogeneous media. The chapter is organized in
two parts: in the first part the theory is developed for linear elastic media, and in the second
part the media is assumed linear viscoelastic. In both cases the properties of the medium are
assumed to be arbitrary (hence not necessarily continuous) functions of the depth y. Explicit
solutions, however, are presented only for the case of a finite number of homogeneous
layers overlaying a homogeneous half-space (a multi-layered medium).
The first topic to be discussed is the Rayleigh eigenvalue problem from which
fundamental results such as the Rayleigh Greens function and the effective Rayleigh phase
velocity can be easily derived. The solution of the Rayleigh eigenvalue problem also leads to
the important concept of geometric dispersion, a phenomenon by which, in heterogeneous
media, the phase velocity of Rayleigh waves is a multi-valued function of the frequency of
excitation. Geometric dispersion needs to be distinguished from material dispersion
introduced in Chapter 2. Whereas the latter is caused by the causality constraint imposed by
the Kramers-Krnig relation, the former arises from constructive interference phenomena
occurring in media that are either bounded (e.g. rods, plates, and other types of waveguides)
or heterogeneous. Geometric dispersion is responsible for the existence of several modes of
propagation each traveling at a different phase and group velocity (modal velocities). Later
in this chapter it will be shown that another consequence of geometric dispersion is to alter
the geometric spreading law governing the attenuation of Rayleigh waves in heterogeneous
elastic media.
For surface waves generated by harmonic forces applied at the boundary or in the
interior of a vertically heterogeneous half space, the various modes of propagation of
Rayleigh waves are superimposed like in a spatial Fourier series. The phase velocity of the
resulting waveform can be obtained from an appropriate superposition of modal Rayleigh
quantities (phase and group velocities, eigenfunctions, etc.). This kinematical quantity is
given the name effective Rayleigh phase velocity and is shown to be a local quantity in the
sense that its magnitude depends on the spatial position where it is measured. It will be
shown in Chapters 5 and 6 that the notion of effective Rayleigh phase velocity is particularly
relevant for surface waves measurements conducted with harmonic sources.

57

58

Rayleigh Waves in Vertically Heterogeneous Media

The last topic to be discussed in this chapter is the Rayleigh variational principle. The
application of this powerful principle to the solution of the Rayleigh eigenvalue problem
leads to closed-form solutions for the partial derivatives of Rayleigh phase velocity with
respect to the medium parameters, in particular the medium compression and shear wave
velocities. As shown in Chapter 5, these partial derivatives are very important in the solution
of the inverse problem in which a given set of Rayleigh phase velocities (i.e., a dispersion
curve) are used to obtain an unknown profile of medium parameters.
The efficiency and accuracy of an inversion algorithm is strongly dependent upon the
technique used to compute the partial derivatives of Rayleigh phase velocity with respect to
the medium parameters. The great advantage offered by the partial derivatives obtained with
the Rayleigh variational principle is that they can be computed using only the unperturbed
medium properties. Numerical partial derivatives, on the other hand, are very inefficient to
compute because they require the solution of the Rayleigh eigenvalue problem for both
perturbed and unperturbed medium parameters. Finally it should be remarked that
numerical computation of partial derivatives is a notoriously ill-conditioned problem. In this
chapter it will be shown how the Rayleigh variational principle can successfully be used to
compute the partial derivatives of both Rayleigh modal and effective phase velocities with
respect to the medium parameters, in a systematic and efficient manner.
Another important result that can be obtained with the use of variational techniques is
related to the attenuation of Rayleigh waves in weakly dissipative media. Some of the most
common procedures used by seismologists to solve surface waves propagation problems in
inelastic media are based on the assumption of weak dissipation (Keilis-Borok, 1989;
Herrmann, 1994). One important consequence of this assumption is that Rayleigh
attenuation coefficients R ( ) can be easily computed from the solution of the Rayleigh
eigenproblem in elastic media. Later in the chapter, a new technique for the solution of the
Rayleigh eigenproblem in linear viscoelastic media is presented. The technique is quite
general since it can also be applied to strongly dissipative viscoelastic media.
The solutions presented in this chapter are all obtained in the frequency domain. This
choice was made for two main reasons. The first reason is simplicity. The mathematics of
wave propagation problems is often fairly involved, and explicit non-integral solutions can
rarely be obtained. One of the few exceptions is a boundary value problem where the
boundary conditions and the body forces are specified as harmonic functions of time. The
second reason for choosing to work in the frequency domain is generality. The availability
of harmonic solutions is often sufficient for obtaining far more general solutions by using
the Fourier integral theorem.
3.2 Rayleigh Eigenvalue Problem in Elastic Media
It is a well established result of classical mechanics (Goldstein, 1980) that the dynamic
behavior of a continuous system can be described by a scalar function, called the Lagrangian

Rayleigh Waves in Vertically Heterogeneous Media

59

density and denoted by L, of a certain number N of generalized coordinates q i ( i = 1, N) and


their spatial and temporal derivatives q i , j and q& i ( j = 1, 3) . The equations governing the
motion of such a system can be derived from Hamiltons principle and are called Lagranges

equations. When the system is conservative, the Lagrangian density L q i , q& i , q i , j is equal
to L = T U where T and U are the kinetic and the potential energy of the system,
respectively. For a conservative system Lagranges equations are (Achenbach, 1984):

L 3 L

+
t q& i j=1 x j q i,j

( )

L = 0
q i

(3.1)

where x j are the components of the position vector x in Cartesian coordinates. For a
linear elastic material the potential energy identifies with the elastic strain energy, and thus
the Lagrangian density is given by:

1
1
L u& i , u i ,j = u& i u& i ij ij
2
2

(3.2)

where u i ( i = 1,3) are the components of the displacement vector, ij and ij are the
components of stress and strain, respectively, and is the mass density that is assumed
constant with time. It is seen from Eq. 3.2 that in a linear elastic body the generalized
coordinates q i identify with the components of displacement vector u ( x, t) .

1
u + u j, i , ij is the
2 i ,j
Kronecker symbol, and and G are Lams elastic moduli, Eq. 3.2 can be rewritten as:
Considering Hookes law ij = kk ij + 2G ij where ij =

1
2
1

L u& i , u i ,j = u& i u& i ( kk ) + G ij ij


2
2

(3.3)

In general, Lams parameters and the mass density are functions of the coordinates, namely

( )

( )

( )

= x j , G = G x j , and = x j . However, in this chapter the elastic medium is

assumed vertically heterogeneous hence = ( x 3 ) , G = G( x 3 ) , and = ( x 3 ) . Using the


definition of Lagrangian density given by Eq. 3.3, Lagranges equations of motion (Eq. 3.1)
yield the following result in vector notation (Ben-Menahem and Singh, 1981):

60

Rayleigh Waves in Vertically Heterogeneous Media


G 2 u + ( + G) grad (div u) + e y

u
d
dG
2 u
e y curl u + 2 = 2
div u +
dy
dy
y
t

(3.4)

Equations 3.4 are the Naviers equations of motion for vertically heterogeneous media
in absence of body forces. Obviously Eq. 3.4 is written in Cartesian coordinates specified by

{}

{ }

a set of basis vectors e j and a set of coordinate axes x j . For convenience the base
vector e 3 has been denoted by e y . Finally, the symbol () ( ) is used to indicate the vector
product.
To find a solution of Eq. 3.4 for harmonic Rayleigh waves, the displacement field
u( x,t ) is assumed equal to:

u: u1 = r1 ( y, k, ) ei( t kr ) , u2 = 0, u3 = i r2 ( y, k, ) ei( t kr )

(3.5)

In elastic media the Rayleigh wave particle motion is elliptical with the minor axis of the
ellipse parallel to the free surface. The horizontal and the vertical components of the
displacement field are 2 radians out of phase. Equation 3.5 represents a two-dimensional
plane strain field ( u 2 = 0 ). This assumption does not imply any loss of generality in the
discussion since it can be proven (Aki and Richards, 1980) that cylindrical Rayleigh waves
have the same y-dependence indicated by Eq. 3.5. In these equations the term k = k( )
denotes the real wavenumber which, in general, is a multi-valued function of the frequency
of excitation . Finally the term r is used to indicate the direction of propagation. Figure
3.1 illustrates the sign convention assumed for the coordinate axes.
To represent Rayleigh waves Eq. 3.5 must be supplemented with appropriate boundary
conditions: no stresses at the free surface of the half-space and no stresses and
displacements at infinity (i.e., the radiation condition):
(r, y) n = 0 at y = 0
u (r , y) 0, (r , y) n 0 as y

(3.6)

where the symbol (r , y ) denotes Cauchys stress tensor and n is a unit normal vector.

Rayleigh Waves in Vertically Heterogeneous Media

Rayleigh Wave

61

n
r
y

Medium Elastic Properties

Figure 3.1

(y ), (y ), G (y )

Rayleigh Waves in Vertically Heterogeneous Media

In vertically heterogeneous media where the material properties ( , G , and ) have


jump discontinuities, the stress and displacement fields must be continuous at each layer
interface:
u (r , y + ) = u (r , y )

( )

(r, y + ) n = r, y - n

(3.7)

If Eq. 3.5 is substituted into the Naviers equations (Eq. 3.4) the result, written in matrix
form, is (Aki and Richards, 1980):

62

Rayleigh Waves in Vertically Heterogeneous Media

r1



r2
d
=
dy
r3



r4

k( ) ( y ) (y ) + 2 G( y )

k ( ) ( y ) 2 (y )

k( )

G 1 ( y )

r1



1
r2
( y ) + 2 G (y )


1
k( ) ( y ) ( y ) + 2 G ( y ) r 3



r4
0
0

2 (y ) k ( )

(3.8)

dr

dr
where r3 ( y , k , ) = G 1 kr2 , and r4 ( y , k , ) = ( + 2 G) 2 + kr1 . The function
dy
dy

( y ) depends on Lams parameters and is given by ( y ) = 4 G

( + G)
.
( + 2G )

The motivation for introducing the functions r3 ( y , k , ) and r4 ( y , k , ) is the following


result:
yr = r3 ( y, k, ) ei(t kr )

(3.9)

yy = i r4 ( y , k, ) ei(t kr )

By defining a vector f ( y ) = r1

r2

r3

r4

, and a matrix A ( y) denoting the 4 4

array above whose elements are functions of ( y), G(y ), k , , Eq. 3.8 can be rewritten
simply as:
df ( y)
dy

= A( y) f ( y)

(3.10)

Equation 3.10 defines a linear differential eigenvalue problem with displacement

eigenfunctions r1 ( y , k , ) and r2 ( y , k , ) and stress eigenfunctions r3 ( y , k , ) and


r4 ( y , k , ) . The boundary conditions associated with the eigenproblem are easily derived
from Eq. 3.6 and Eq. 3.9:

Rayleigh Waves in Vertically Heterogeneous Media


r3 (y, k, ) = 0,

63

r4 ( y, k, ) = 0 at y = 0

(3.11)

f (y , k, ) 0 as y

For a given frequency , non-trivial solutions of the linear eigenproblem (Eq. 3.10)
subjected to the boundary conditions (Eq. 3.11) exist only for special values of the
wavenumber k j = k j ( ) , ( j = 1, M) . These particular values of k j are the eigenvalues of

) (i = 1, 4)

the eigenproblem, and the corresponding solutions ri y , k j , ,

are the

eigenfunctions. It can be shown (Keilis-Borok, 1989) that the set of eigenfunctions for a
given frequency obey certain orthogonality conditions with appropriate weighting
functions.
The

relation

k j = k j ( )

is

only

known

in

the

implicit

form

F R ( y ), G( y ), ( y ), k j , = 0 where F R [] is a complicated function of Lams


parameters, the mass density, the wavenumber and the frequency of excitation. The
relationship F R [] = 0 is called the Rayleigh dispersion equation. It states that in vertically
heterogeneous media the velocity of propagation of Rayleigh waves is, in general, a multi-

valued function of frequency. Each pair k j , ri y , k j ,

)} defines a mode of propagation

and, in general, there are M normal modes of propagation at any given frequency. The
number M can be finite or infinite, depending on the y-dependence of the medium
properties and on the frequency of excitation. Furthermore, the distribution of the modes,
namely the mode spectrum, can be continuous or discrete, and in some cases both (KeilisBorok, 1989). In a medium composed of a finite number of homogeneous layers overlaying
a homogeneous half-space, the total number of Rayleigh modes of propagation is always
finite (Ewing et al., 1957).
From a physical point of view, the existence of different modes of propagation at a
given frequency can be explained by the constructive interference occurring among waves.
In continuously varying heterogeneous media the ray paths are curved (as a result of
Fermats principle), and hence they interfere with each other. In multi-layered media the ray
paths are rectilinear, and interference phenomena occur among waves undergoing multiple
reflections at the layer interfaces. In either case, the dispersion equation is the mathematical
statement of this condition of constructive interference (Achenbach, 1984).
3.2.1

Solution Techniques

Several techniques may be used to solve the linear eigenvalue problem with variable
coefficients including numerical integration, finite difference, finite element, boundary
element, and spectral element methods. However, some of the most common and

64

Rayleigh Waves in Vertically Heterogeneous Media

frequently used techniques are those belonging to the class of propagator-matrix methods
(Kennett, 1983). The Thomson-Haskell algorithm (Thomson, 1950; Haskell, 1953) also
called the transfer matrix method is probably the most famous of this class of methods
because of its conceptual simplicity and ease of computer implementation. The application
of this algorithm, however, is limited to vertically heterogeneous media that can be
represented by a stack of homogeneous layers overlying a homogeneous half-space.
are

In the Thomson-Haskell algorithm the non-trivial solutions of the linear eigenproblem


found
from
the
roots
of
the
Rayleigh
dispersion
equation

F R ( y ), G( y ), ( y ), k j , = 0 . The dispersion equation is constructed by a sequence of

matrix multiplications involving terms that are transcendental functions of the material
properties of the layers in the stratified medium. The roots of the Rayleigh dispersion
equation are the wavenumbers corresponding to the modes of propagation of Rayleigh
waves at each frequency. Once the roots of the Rayleigh dispersion equation are found, the
eigenfunctions for each mode of propagation can be computed by means of simple
algebraic manipulations. The Rayleigh eigenfunctions give the depth-dependence of the
stress and displacement.
Haskell (1953) also developed asymptotic expressions for the Rayleigh dispersion
equation in the important limiting cases of short and long wavelengths. Because the original
formulation of the Thomson-Haskell algorithm suffers numerical instability problems at
high frequencies (Knopoff, 1964), this method has been modified and improved throughout
the years by numerous researchers (Schwab and Knopoff, 1970; Abo-Zena, 1979; Harvey,
1981).
Kausel and Rosset (1981) derived a finite element formulation from the ThomsonHaskell algorithm, which is called the dynamic stiffness matrix method. The main feature of
this method is the replacement of the Thomson-Haskell transfer matrices with layer
stiffness matrices that are similar to conventional stiffness matrices used in structural
analysis. The advantage of this formulation is the ability to use standard structural analysis
techniques such as condensation and substructuring to solve both the eigenproblem and the
inhomogeneous elastodynamic problem of layered media subjected to dynamic loads
(Kausel, 1981). The first attempts in using finite element techniques to solve wave
propagation problems in seismology and earthquake engineering date back to the early 70s
with the works of Lysmer and Waas (1972) and Lysmer and Drake (1972).
Another important class of algorithms for solving eigenvalue problems of surface waves
is the method of reflection and transmission coefficients developed by Kennett and his coworkers (Kennett, 1974; Kennett and Kerry, 1979; Kennett, 1983), and modified and/or
improved by others researchers (Luco and Apsel, 1983; Chen, 1993; Hisada, 1994; Hisada,
1995). This method, like the Thomson-Haskell algorithm to which it is related, is only
suitable for applications in multi-layered media. It is based on the use of reflection and
transmission coefficients to construct reflection and transmission matrices for a stratified

Rayleigh Waves in Vertically Heterogeneous Media

65

media. The result is a very efficient iterative algorithm for establishing the Rayleigh

dispersion equation F R ( y ), G( y ), ( y ), k j , = 0 . The method of reflection and

transmission coefficients also offers an interesting physical interpretation because it


explicitly models the constructive interference that leads to formation of the surface waves
modes (Kennett, 1983). Earlier versions of this algorithm were numerically unstable at high
frequencies because of the presence of certain frequency-dependent terms that have been
eliminated in more recent formulations (Chen, 1993).
Most of the computational efforts spent by the algorithms used for the solution of the
Rayleigh eigenvalue problem are devoted to the following two tasks: construction of the

Rayleigh dispersion equation F R ( y ), G( y ), ( y ), k j , = 0 (which is also called the


Rayleigh secular function), and computation of its roots as a function of frequency. The
latter are the Rayleigh wavenumbers k j = k j ( ) , and also the eigenvalues solution of the
Rayleigh eigenproblem (Eqs.3.10 and 3.11).
In the case of an elastic medium the use of complex arithmetic in constructing the
Rayleigh secular function can completely be avoided (Haskell, 1953; Schwab and Knopoff,
1971), and the roots of the dispersion equation are generally obtained by means of rootbracketing techniques combined with bisection (Hisada, 1995). The use of these slow
converging root-finding techniques is suggested by the rapidly oscillating behavior of the
Rayleigh secular function, particularly at high frequencies, which requires the use of methods
that cannot fail to find the roots (Press et al., 1992).
Figure 3.2 shows a typical plot of the roots of the Rayleigh dispersion equation
k j = k j ( ) where the phase velocity, rather than the wavenumber, has been plotted against
frequency. The algorithm used is that developed by Chen (1993) and Hisada (1995) which
uses the method of reflection and transmission coefficients. Each dispersion curve is
associated with a particular mode of propagation.
In general, there are several modes of propagation at a given frequency with the higher
modes characterized by a higher velocity of propagation. Another important feature of
multi-mode Rayleigh wave propagation is readily apparent from Fig. 3.2: as the frequency
increases the number of modes associated with that frequency also increases and the modes
become more closely spaced. As the modes all tend to a common limit which is
the Rayleigh phase velocity of the thin layer bordering the free-surface of the vertically
heterogeneous half-space.
Once the roots of the Rayleigh secular function have been obtained, computation of the
eigenfunctions is a straightforward task. Figure 3.3 illustrates the typical mode shapes of
Rayleigh displacement eigenfunctions in a vertically heterogeneous medium.

66

Rayleigh Waves in Vertically Heterogeneous Media


165

Rayleigh Phase Velocity (m/sec)

160

155

150

3
145

2
140

135

1
130

125
0

20

40

60

80

100

Frequency (Hz)

Figure 3.2

Rayleigh Waves Dispersion Curves in Vertically Heterogeneous Media


Mode Number
1

-5

Vertical

Depth (m)

-10

Horizontal
-15

-20

-25

-0.5

0.5

1.5

2.5

Normalized Particle Displacement

Figure 3.3

Rayleigh Displacement Eigenfunctions in Vertically Heterogeneous Media

Rayleigh Waves in Vertically Heterogeneous Media

67

The eigenfunctions have been normalized with respect to the maximum value of the
vertical displacement. A common feature of multi-mode Rayleigh wave propagation is that
higher modes have a greater penetration depth than lower modes. This property, which is
clearly shown in Fig. 3.3, is very important in the solution of the inverse problem because
the resolution of deeper layers can be directly related to the presence of higher modes of
propagation.
As mentioned at the beginning of this section, other numerical techniques can be used
to solve the linear eigenvalue problem including the finite difference method (Boore, 1972),
numerical integration (Takeuchi and Saito, 1972), the boundary element method (Manolis
and Beskos, 1988), and the spectral element method (Komatitsch and Vilotte, 1998; Faccioli
et al., 1996). Although these techniques are less popular than the propagator matrix
methods briefly described in this section, they have several advantages. Numerical
integration and spectral element method, for example, can be used in vertically
heterogeneous media where the medium properties vary continuously with depth, and
therefore they are more general than the propagator matrix methods. Boundary element
methods are best suited for modeling unbounded or semi-infinite media because they
require discretization of only the boundaries. As a result, they reduce the dimension of the
problem by one. Moreover, boundary element methods eliminate the need, required by
finite element based methods, of using fictitious or non-reflecting boundaries to simulate
the radiation condition at infinity.
3.3 Effective Rayleigh Phase Velocity in Elastic Media
In Section 3.2 it was shown that several Rayleigh wave modes may propagate in a
vertically heterogeneous medium at a given frequency . Each mode is specified by the pair

{k , r (y , k ,)} where k ()
j

(j = 1, M)

is the wavenumber, and ri y , k j ,

) (i = 1, 4) is

the set of four eigenfunctions. M = M( ) is the total number of modes associated with the
frequency . The linear eigenvalue problem corresponds formally to a homogeneous
boundary value problem, and its solutions are known as free Rayleigh waves (Ewing et al.,
1957). In inhomogeneous problems, Rayleigh waves are generated by sources applied at the
boundary or in the interior of the half-space. If these sources are harmonic in time, the
different Rayleigh modes of propagation are superimposed on one another. The phase
velocity of the resulting wave train is here named the effective Rayleigh phase velocity. This
section is devoted to deriving an explicit, analytical expression for this kinematical quantity
from the solution of the Rayleigh eigenproblem.
In isotropic vertically heterogeneous media, Rayleigh waves generated from point
sources acting in a direction perpendicular to the boundary of the half-space propagate
along cylindrical wave fronts (Ewing et al., 1957). It can be shown (Ben-Menahem and
Singh, 1981) that the wave field originated by an harmonic point source located at a position

68

Rayleigh Waves in Vertically Heterogeneous Media

(r = 0, y = y )

can be expanded, in the radial direction, in a series of pth order Hankel


functions (p is an integer). For large values of r the pth order Hankel functions can be
approximated by their asymptotic expansions involving complex exponential functions. As
S

a result, the particle displacement u (r , y , ) = u r (r , y , ) e r + u y (r , y , ) e y resulting from

the superposition of M distinct Rayleigh modes, can be written in cylindrical coordinates

{r , y , } , as follows (Aki and Richards, 1980):


u (r , y , ) =

[ A ( r , y , )] e (
M

i t k j r +

j= 1

(3.12)

where = r , y and A (r , y , ) , k j ( ) are the Rayleigh displacement amplitudes and


j

wavenumber, respectively, associated with the jth mode of propagation. Finally, = 4

for = r and = 4 for = y . Equation 3.12 shows, as expected, that u (r , y , ) is


independent of the azimuthal angle . The actual particle displacement is obtained by
taking either the real or imaginary part of Eq. 3.12. By choosing the latter, Eq. 3.12
becomes:
M
M
i (t k j r + )

u (r , y , ) = A (r , y , ) e
= C j sin (t) D j cos (t)
j

j=1
j=1

where

(C ) = (A ) cos(k r + )

( )

and

( )

(D ) = (A ) sin (k r + ) .

(3.13)

Now using simple

trigonometric identities Eq. 3.13 can be re-written as follows:

u (r , y , ) =U (r , y , ) sin t (r , y , )

(3.14)

where:

[ (

M M
U (r , y , ) = A (r , y , ) A (r , y, ) cos r k i k j
i
j
i=1 j=1

] [

)]

0.5

(3.15)

Rayleigh Waves in Vertically Heterogeneous Media


M
A (r , y , ) i sin k i r +
1 i =1
(r , y , ) = tan M
A (r , y , ) cos k r +
j

j
j=1

[
[

]
]

(
(

69

(3.16)

From Eq. 3.14, the expression:

[t (r , y , )] = cons tan t

(3.17)

represents the equation of a wave front, since it is the locus of points having constant

phase. Assuming the functions (r , y , ) to be sufficiently smooth, Eq. 3.17 can be


differentiated with respect to time, to give:

=0
( r, y, ) dr
dt

(3.18)

which finally yields:


V$ (r , y , ) =

[ (r , y , )]

(3.19)
,r

where the symbol V$ (r , y , ) has been used to denote the effective Rayleigh phase velocity.
It is apparent from Eq. 3.19 that the effective Rayleigh phase velocity is a local quantity,
which means that its value depends on the spatial position where it is evaluated. At a fixed
y = y , the function V$ (r , y c , ) describes what may be called a dispersion surface, i.e. a two
c

dimensional surface showing the variation of the effective Rayleigh phase velocity with
frequency and distance from the source. From Eq. 3.19 it is also interesting to observe that
since the effective Rayleigh phase velocity is a vector quantity, different components of
V$ (r , y , ) will, in general, travel at different phase velocities. Furthermore, since
V$
t

( )
( ) ( )
V$

,r

,r

is not, in general, equal to zero, the wave train accelerates as it


,r

70

Rayleigh Waves in Vertically Heterogeneous Media

propagates along the surface of the half-space. In Eq. 3.19 the term

( )

,r

could be

interpreted as an effective Rayleigh wavenumber and denoted as k$ (r , y , ) . However, a

decomposition of the argument of Eq. 3.14 in a form t k$ r , which is common for

( )

harmonic waves, is no longer possible here because the effective wavenumber

,r

being

a local quantity, must be integrated over r to yield the phase ( r , y , ) . Considering the

definition of (r , y , ) given by Eq. 3.16, it is possible to obtain from Eq. 3.19, an explicit
definition for the effective Rayleigh phase velocity which is given by:

( A
M

V$ ( r , y , ) = 2

i =1 j = 1

{( A
M

r = 1 s =1

) (A ) cos [r ( k
i

) (A ) (k
r

)]

k j

+ k s ) cos r ( k r k s )

(3.20)

]}

For an harmonic point source Fy e it located at r = 0, y = y S , the Rayleigh displacement


amplitudes

[A (r , y , )]

of the individual modes of propagation are related to the

displacement eigenfunctions r1 ( y , k , ) and r2 ( y , k , ) , and to other modal parameters by


the following expression (Aki and Richards, 1980):

A r ( r , y , )

Fy r2 y S , k j ,
=
A (r , y , ) =
j
4 Vj U j I j 2 r k j

A y ( r , y , ) j

r y , k ,
j
1

r y , k ,
j
2

)
)

(3.21)

where Vj , U j , and k j are the phase, group velocity and wavenumber of the Rayleigh jth

mode of propagation ( j = 1, M) , respectively. The term I j y , k j , is the first Rayleigh


th

energy integral associated with the j mode of propagation (see Section 3.5 for more details)
and is defined by (Aki and Richards, 1980):

[ (

) (

)]

1
I j ( y , k j , ) = ( y ) r12 y , k j , + r22 y , k j , dy
20

(3.22)

Rayleigh Waves in Vertically Heterogeneous Media

71

By substituting Eq. 3.21 into Eq. 3.20, the expression for the effective Rayleigh phase
velocity becomes:

( )

) [(

)]

r ( y , k )r y , k r ( y , k )r y , k cos r k k
i 1
j 2
S
i 2
S
j
i
j
1

i = 1 j =1
Vi U i I i ) Vj U jI j k i k j
(

V$ r (r , y , ) =
M M r ( y , k )r ( y , k )r ( y , k )r ( y , k )( k + k ) cos r ( k k )
2
1
r 1
s 2
S
r
S
s
r
s
r
s

( Vr U r I r )( Vs U s Is ) k r k s
r =1 s = 1

( )

) [(

(3.23a)

)]

r (y , k )r y , k r (y , k )r y , k cos r k k
i 2
j 2
S
i 2
S
j
i
j
2

i = 1 j =1
V i U i I i ) Vj U j I j k i k j
(

V$ y (r , y , ) =
M M r ( y , k )r ( y , k )r ( y , k )r ( y , k )( k + k ) cos r ( k k )
r
s 2
S
r
S
s
r
s
r
s
2
2
2

(Vr U r I r )( Vs U s Is ) k r k s
r =1 s = 1

(3.23b)

where V$ r (r , y , ) and V$ y (r , y , ) denote the components of the effective Rayleigh phase


velocity along directions r and y respectively. To reduce the length of the above expressions,

the frequency dependence of the eigenfunctions r1 ( y , k , ) and r2 ( y , k , ) has been


omitted.

As a final remark of this section, it is noted from Eq. 3.23 that the effective Rayleigh
phase velocity is completely determined by the solution of the Rayleigh eigenproblem. In
fact, recalling that Vj = k j and U j = d dk j

(j = 1, M) ,

all of the modal quantities

appearing in Eq. 3.23 can be calculated from the pair k j , ri y , k j ,

)} (i = 1, 2) .

3.4 Rayleigh Greens Function in Elastic Media


The term Greens function is generally used in applied mathematics to denote the
response of a linear system governed by a set of differential or integral equations and
associated boundary conditions (integral equations contain built-in boundary conditions) to
an impulsive unit point source represented by a Dirac- ( ) distribution. It is possible to
show (Logan, 1997) that the response of a system to an arbitrary distribution of sources in
space and time can uniquely be determined by the knowledge of the Greens function of the
system via a convolution integral. In engineering, the term Greens function is frequently

72

Rayleigh Waves in Vertically Heterogeneous Media

used with a somewhat more general and sometimes different meaning than the one used in
applied mathematics. A common example is constituted by the response of a linear system
to a harmonic or a Heaviside unit point source. This response if often denoted as the
Greens function associated to the linear system.
In this study the term Greens function is primarily used to denote the response of a
linear elastic or viscoelastic half-space to a harmonic unit point source. Of particular interest
in this section is the displacement Greens function, which is defined as the displacement
u$ (r , y , ) induced in a linear elastic medium by a harmonic unit point load 1 y e it located
at the position (r = 0, y = y S ) . The subscript y in 1 y denotes the direction of action of the
unit point load.
The particle displacement field u$ (r , y , ) can be separated into two components

u$ (r , y , ) = u$ B (r , y , ) + u$ S (r , y , ) . The first component u$ B (r , y , ) represents the body


wave field and is composed of the superposition of P and S waves. The component
u$ S (r , y , ) is the surface wave field and, in general, is composed of a superposition of Love
and Rayleigh waves. The body wave field attenuates with distance from the source at a
much higher rate than the surface wave field. It is possible to show (Ewing et. al, 1957) that
at the free surface of a homogeneous half-space, body waves generated by harmonic point
sources attenuate with a factor proportional to r 2 . For the same setting, the spatial
attenuation of surface waves is proportional to r 0.5 . These laws of attenuation are not
applicable, in general, to transient wave-forms because in the latter, the spatial attenuation
of the wave results from a combination of both geometrical spreading and spreading of the
signal with time (Keilis-Borok, 1989).
It is evident from these considerations that for harmonic oscillations at large distances
from the source, the surface wave field dominates the overall particle motion and
u$ (r , y , ) u$ S (r , y , ) . The distance from the source where the body wave field is not
negligible is usually called the near field. Numerical studies by Holzlohner (1980), Vrettos
(1991) and Tokimatsu (1995) of wave propagation in vertically heterogeneous media have
shown that in normally dispersive media the near-field effects are important up to a distance
from the source equal to 2 (where = ( ) is the wavelength of the Rayleigh waves).
However, in inversely dispersive media (i.e. media where the material properties vary
irregularly with depth) the near-field is larger and may extend up to 2 . All theoretical
studies and experimental measurements performed during this research program have
assumed the near-field effects to be negligible. Furthermore, even though surface waves
tests are suitable for measurements of both Love and Rayleigh waves, most current
applications including this study focus exclusively on Rayleigh waves.

Rayleigh Waves in Vertically Heterogeneous Media

73

With these assumptions the displacement Greens function u$ (r , y , ) can be easily


computed from the solution of the Rayleigh eigenproblem using the concept of mode
superposition. In the previous section it was shown how to compute the particle

displacement induced by a harmonic point source Fy e it located at r = 0, y = y S of a


vertically heterogeneous half-space. From Eq. 3.14,
i t ( r , y , ) ]
u$ (r , y , ) =U$ (r , y , ) e [

(3.24)

where the subscript = r , y denotes the radial and the vertical directions, respectively.
The expressions for U$ (r , y , ) and (r , y , ) are obtained from Eqs. 3.15 and 3.16

with the modal amplitudes A (r , y , ) computed from Eq. 3.21 with Fy = 1; the final
j

result is:

) [(

)]

) [(

)]

M M r ( k , y )r k , y r ( k , y )r k , y cos r k k
i
j
i
S
j
S
i
j
1
1
2
2
1

$
U r (r , y , ) =

4 2 r i = 1 j= 1
k i k j ( V i U i I i )( V j U j I j )

M M r ( k , y )r k , y r ( k , y )r k , y cos r k k
i
j
i
S
j
S
i
j
2
2
2
2
1

$
U y (r , y , ) =

4 2 r i =1 j=1
k i k j ( V i U i I i )( V j U j I j )

0 .5

(3.25a)

0 .5

(3.25b)

and
M r ( k , y , )r ( k , y , )

2
i
S
1 i
sin k i r

4
i =1
k i ( Vi U i I i )
r ( r , y , ) = tan 1 M

r1 ( k j , y , )r2 ( k j , y S , )

cos k j r

4
j =1
k

V
U
I
j
j j j

(3.26a)

74

Rayleigh Waves in Vertically Heterogeneous Media

M r ( k , y , )r ( k , y , )

2
i
S
2 i
sin k i r +

4
i =1
k i ( Vi U i I i )
y (r , y , ) = tan 1 M

r2 ( k j , y , )r2 ( k j , y S , )

cos k j r +

4
j =1
k

V
U
I
j
j
j
j

(3.26b)

Equations 3.24 through 3.26 completely define the displacement Greens function

u$ (r , y , ) . Equation 3.24 is informative because it shows that a multiplicative

i arg ( u$ )
decomposition of the displacement Greens function of the type u$ =U$ e
is possible

even for multi-mode wave propagation. However, because the wavenumber k$ (r , y , ) is


no longer a constant, the spatial variation of the displacement field is no longer harmonic
even though the temporal variation of the source is harmonic. Equations 3.25 and 3.26
show also the remarkable result that the three main factors in the expression of u$ (r , y , ) ,
namely the source depth ( y S ) , the receiver depth ( y ) , and the distance from the source
( r ) , are uncoupled in the sense that their contribution is independent from each other.

With the definition of Greens function given in this section, the computation of the

displacement u (r , y , ) induced by a point harmonic source Fy e it located at

{r = 0, y = y } becomes a trivial task. In fact:


S

u (r , y , ) = Fy u$ (r , y , )

(3.27)

It is instructive to re-write Eq. 3.27 as follows:


u (r , y , ) = Fy G (r , y , ) e [

i t ( r ,y , )

(3.28)

where a new function G (r , y , ) =U$ (r , y , ) called the Rayleigh geometrical spreading function,
has been introduced. This function has the important physical interpretation of modeling
the geometric attenuation in vertically heterogeneous media. As already mentioned Rayleigh
waves in homogeneous media attenuate by a factor proportional to r 0.5 as a result of their
geometrical spreading from a localized source. This simple geometric attenuation law, which
follows directly from the principle of conservation of energy, does not hold in non-

Rayleigh Waves in Vertically Heterogeneous Media

75

homogeneous media. Under these conditions the Rayleigh wave displacement field results
from the superposition of several modes of propagation (geometric dispersion). An
important consequence is that the geometric spreading law of Rayleigh waves is altered.
Figure 3.4 shows a plot of the geometric spreading function Gy (r , y , ) at the free-

surface ( y = 0) of three different types of elastic media: a homogeneous medium, a threelayer soft-stiff-stiffer system (normally dispersive), and a three-layer stiff-soft-stiff system
(inversely dispersive). The numerical simulation was carried out at a frequency of 40 Hz.

1.0E-08

Homogeneous

Gy (r,y,) [m]

8.0E-09

Inversely Dispersive

6.0E-09

Normally Dispersive
4.0E-09
2.0E-09
0.0E+00
0

10

20

30

40

50

60

Distance [m]
Figure 3.4

Geometric Spreading Function for Different Types of Media

It is apparent from the figure that geometric attenuation in inversely dispersive media is
most strongly affected by geometric dispersion. This has also been observed in other
numerical studies (Tokimatsu et al., 1992; Gucunski and Woods, 1991).
From Eq. 3.25 it can be easily verified that in homogeneous media where M = 1 ,

k i = k j = k , and U i = Vi = V , G (r , y , ) reduces as expected, to the frequency


independent

function

where

E r = r1 ( y ) r2 ( y S ) 4 V 2 I 2 k

and

E y = r2 ( y ) r2 ( y S ) 4 V 2 I 2 k . The importance of the explicit factorization of the

76

Rayleigh Waves in Vertically Heterogeneous Media

Rayleigh displacement field into a product of source magnitude, geometric spreading


function, and phase factor will become apparent in Section 5.3, which is dedicated to the
attenuation of Rayleigh waves in weakly attenuating media.
3.5 Variational Principle of Rayleigh Waves in Elastic Media
This section will illustrate some interesting results that can be obtained from the
application of Hamiltons principle to the solution of the Rayleigh eigenvalue problem. The
most important of these results will be the derivation of closed-form expressions for the
partial derivatives of Rayleigh phase velocity with respect to the body wave velocities of the
medium. As already mentioned in Section 3.1, these partial derivatives are very useful in the
solution of the inverse problem, a topic that will be discussed in Chapter 4. The Rayleigh
variational principle can be used to compute both the modal and the effective Rayleigh
phase velocity partial derivatives with respect to the medium parameters. The Rayleigh
variational principle can also be used to obtain important results for Rayleigh waves
propagating in weakly attenuating media. In fact, it will be shown later in this section that
the solution of the Rayleigh eigenproblem in elastic media can be used as a basis to compute
the Rayleigh attenuation coefficient R ( ) , which is a parameter characterizing the
response of dissipative media.
Hamiltons principle applied to a continuous, conservative system of volume V, states
that among all the possible paths of motion between two instants in time t 1 and t 2 the
actual path is such that the integral:

t2

I = L qi , q& i , qi,j dVdt


t1

(3.29)

where L = T U has a stationary value (Goldstein, 1980). The function L q i , q& i , q i , j

( i = 1, N; j = 1, 3) , already introduced in Section 3.2, is called the Lagrangian density,


whereas the variables q i describing the behavior of the system are the generalized
coordinates. Finally T and U are the kinetic and the potential energy of the system,
respectively. It is well known from the calculus of variations (Logan, 1997) that for the
integral of Eq. 3.29 to have a stationary value it is required that:
t2

(I ) = L q i , q& i , q i ,j dVdt = 0
t1

(3.30)

Rayleigh Waves in Vertically Heterogeneous Media

77

That is, the first variation of the integral I vanishes for arbitrary changes q i which
vanish at the boundary of the volume V and at times t 1 and t 2 . Implementation of Eq.
3.30 yields Lagranges equations of motion of the system and the associated natural
boundary conditions. It was shown in Section 3.2 that in a continuous linear elastic material

the Lagrangian density is given by L u& i , u i ,j = u& i u& i ( kk ) + G ij ij where


2
2

= ( y ) , G = G( y ) , and = ( y ) . If the elastic body of volume V is identified with a


vertically heterogeneous half space with no body forces and surface tractions, and if the
displacement field u( x,t ) is specified according to Eq. 3.5, the expression for the
Lagrangian density becomes:
2
2
2

dr1

dr2
dr2
1 2 2 2 1
2 2

L = (r1 + r2 ) kr1 + + G kr2 + 2G k r1 +


4
4
dy
dy

dy

(3.31)

where the symbol denotes the average value of a quantity. The average value has been
used to eliminate the time dependence from the definition of the harmonic Lagrangian
density. Equation 3.31 defines the average Lagrangian density L for the homogeneous
boundary value problem of Rayleigh waves. Application of Hamiltons principle with the
Lagrangian density given by Eq. 3.31 for any perturbation of the eigenfunctions r1 ( y , k , )

and r2 ( y , k , ) satisfying the boundary conditions given by Eq. 3.11 yields:

(I ) = L dy = 2 I1 k 2 I2 kI3 I 4 = 0
0

(3.32)

where I1 , I 2 , I3 , and I 4 are the Rayleigh energy integrals, and are defined as (Aki and
Richards, 1980):

I1 =

1
(r12 + r22 )dy
2 0

I3 =

dr2
dr1
r1
dy
Gr2
dy
dy

I2 =

1
( + 2G) r12 + Gr22 dy
2 0

2
2
dr2
dr1
1
I4 = ( + 2G) + G dy
2 0
dy
dy

(3.33a)

(3.33b)

78

Rayleigh Waves in Vertically Heterogeneous Media

In Eq. 3.32 the eigenfunctions r1 ( y , k , ) and r2 ( y , k , ) are the only quantities being
perturbed, hence:

(I ) = L dy = ( 2 I1 k 2 I2 k I 3 I4 ) = 0

(3.34)

From the Rayleigh equations of motion and associated boundary conditions (Eqs. 3.10
and 3.11), it is possible to obtain the following result at the stationary point where

L dy = 0 (Aki and Richards, 1980):


0

L dy = ( 2 I1 k 2 I 2 kI3 I4 ) = 0

(3.35)

Equation 3.35, written in the form I1 = k I 2 + kI3 + I 4 , can be interpreted as a


statement of conservation of energy, i.e., the average kinetic energy associated with a given
mode of propagation equals the average elastic strain energy.
2

The combined results given by Eqs. 3.34 and 3.35 will be referenced hereafter as the
variational principle of Rayleigh waves. This should not be confused with the Rayleigh
principle which asserts that first-order perturbations in the Rayleigh eigenvalue (namely the
wavenumber k) will only result in second-order perturbations of the corresponding
eigenfunctions (Ben-Menahem and Singh, 1981) and may, in fact, be derived from Eqs. 3.34
and 3.35 if desired.
3.5.1

Modal Rayleigh Phase Velocity Partial Derivatives

The results obtained in the previous section, namely, the variational principle of
Rayleigh waves, will now be used to obtain closed-form expressions for the partial
derivatives of the modal Rayleigh phase velocity VR ( ) with respect to the body wave

velocities of the medium VP and VS . For this purpose, let the triple ( y), G( y), ( y )
characterize the material properties of a linear elastic vertically heterogeneous medium M .

The pair k j , ri , G , , k j ,

)} with i = 1, 4; j = 1, M will represent the solution of the

Rayleigh eigenvalue problem associated with this medium. In the functional dependence of
the eigenfunctions, the material properties ( y) , G( y ) , and ( y) have been considered as

[ ]

independent variables. Now let ( y) + ( y ) , G( y ) + G( y ) , ( y ) denote the


~
material properties of a medium M whose Lams parameters differ slightly from those of
the
medium
At
a
given
frequency
,
the
pair
M.

Rayleigh Waves in Vertically Heterogeneous Media

{k + k , r ( + , G + G, , k + k , )}
j

79

represents the solution of the Rayleigh

eigenproblem associated with this new medium.


Later in this section it will be shown that the problem of determining the partial
derivatives of the modal Rayleigh phase velocity VR ( ) j with respect to the medium

parameters is essentially reduced to that of computing k j . The latter task is accomplished


by using the variational principle of Rayleigh waves, namely Eqs. 3.34 and 3.35. For ease of
notation, the modal parameters k j and VR ( ) j will subsequently be denoted without the
~
subscript j. Application of Eq. 3.35 to the medium M yields:

2
2
1 2
(r1 + r1 ) + (r2 + r2 ) dy =
2 0

{[

1
2
2
( k + k) 2 ( + ) + 2 ( G + G) (r1 + r1 ) + ( G + G)(r2 + r2 ) dy +
2
0

d
d
+ ( k + k) ( + )(r1 + r1 ) (r2 + r2 ) ( G + G)(r2 + r2 ) (r1 + r1 ) dy +
dy
dy

1
+ ( + ) + 2 ( G + G)
2 0

(
)
+
+
+
r

r
G

G
r
+
r
(
)
(
)

2
2
1
1
dy

dy

(3.36)

dy

Application of Eqs. 3.34 and 3.35 to Eq. 3.36 expanded to include first order terms leads
to:

80

Rayleigh Waves in Vertically Heterogeneous Media

dr

dr
k r1 2 Gr2 1 + k ( + 2G)r12 + Gr22 dy +
dy
dy

1 dr 2

dr2 1 2 2
2
(
)
+ k r + 2G dy +
+ + kr1
dy 2 1
2 dy

(3.37)

1 dr 2
dr1
dr2 1 2 2
1
+ 2r1 + k r2 Gdy = 0
+ k r2
dy
dy
dy 2
2

Now from Eq. 3.35:

( 2 I1 k 2 I2 kI3 I4 ) =
= 2 + I1 2I 2 kk k I2 I 3 k kI3 I4 = 0
2

(3.38)

which considering Eq. 3.34 yields:

U=

2 kI 2 I3
=
k
2I1

(3.39)

Equation 3.39 is an important result because it is an alternative procedure to compute the


group velocity without using numerical differentiation. In view of Eq. 3.39, Eq. 3.37 can be
re-written as:

1
k =

4UI1 0

2
2

dr1
dr2
dr2
(
)
kr2 4 kr1
Gdy + kr1 + + 2G dy
dy
dy
dy

(3.40)

k2
Since VR = , it follows that k = VR , and hence from Eq. 3.40:
k

1
VR = 2

4 k UI1 0

2
2

dr1
dr2
dr2
kr2 4 kr1
Gdy + kr1 + ( + 2G)dy
dy
dy
dy

(3.41)

Rayleigh Waves in Vertically Heterogeneous Media

81

Equation 3.41 allows one to compute the change in Rayleigh phase velocity VR

resulting from a small perturbation of the Lams parameters ( y) , and G( y ) . Equation


3.41 may also be expressed as:

VR
VR =
G dy +
0 G ,M

VR

M
dy

,G

(3.42)

where M = ( + 2G) is the constrained modulus and:


2

dr1
dr2
1
VR

4
kr2
kr1
2
G
dy
dy
,M 4 k UI1

dr2
1
VR

+
kr
2
1
M
dy
,G 4 k UI1

(3.43)

The subscripts outside the brackets indicate the parameters that are held constant.
Equations 3.43 are known in seismology as the partial derivatives of the modal Rayleigh
phase velocity VR ( ) with respect to the medium parameters G and M . However, the

term partial derivative must be used with care because VR = VR M(y ), G( y), ( y ), k ,

and hence VR ( ) is a functional rather than a function of the parameters M( y ) and G( y ) .


Accordingly, the partial derivatives in Eq. 3.43 are understood to refer to a particular depth
y.
One more step is required to compute the partial derivatives of VR ( ) with respect to
the body wave velocities of the medium VP and VS . Since G = VS2 and M = VP2 , using
the chain-rule of calculus:
2
VR
G
dr1
dr2
VR
VS
VR

=
y
,
k
,
=

kr

4
kr

(
)

2
2
1
G
dy
dy

,M VS 2k UI1
VS ,VP VS

VR
M
V
VP
dr
V
= R ( y , k , ) = R
kr1 + 2

= 2
dy
M ,G VP 2k UI1
VP ,VS VP

(3.44)

82

Rayleigh Waves in Vertically Heterogeneous Media

Equations 3.44 give a measure of the sensitivity of modal Rayleigh phase velocity
VR ( ) to small changes of the medium parameters VP and VS at a specific depth y. The
remarkable feature of Eq. 3.44, which makes it so valuable in the solution of the Rayleigh
inverse problem, is that these partial derivatives can be computed using Rayleigh wave
parameters referred to the original and not the perturbed VP and VS profiles. Conversely,
it would be very expensive to compute the above partial derivatives numerically with, say, a
four-point central finite difference scheme (Spang, 1995); a single computation of
VR VS would require the solution of four Rayleigh eigenproblems instead of just one
eigenproblem using the variational approach.
Examining Eq. 3.44, it is found that the phase velocity of Rayleigh waves is relatively
insensitive to changes in VP (Lee and Solomon, 1979; Ben-Menahem and Singh, 1981), and
thus the partial derivative VR VP is small compared to VR VS . Figure 3.5 shows the
partial derivatives given by Eq. 3.44 for the case of a homogeneous medium and a frequency
of 40 Hz.

dVR/dVS , dVR/dVP
0.0

0.1

0.2

0.3

0.4

0.5

0.0

Depth [m]

2.0
4.0

/d

/d

dVR/dVS
dVR/dVP

6.0
8.0

Frequency 40 Hz
VS = 120 m/s
VP = 400 m/s

10.0
Figure 3.5

Partial Derivatives of Rayleigh Phase Velocity with Respect to VP and VS for


a Homogeneous Medium

Rayleigh Waves in Vertically Heterogeneous Media

83

It is apparent from the figure that the shear wave velocity VS controls the Rayleigh
wave velocity VR ( ) . The largest values of VR VS and VR VP are 0.418 and 0.105
respectively. Both maxima occur at the free-surface (y = 0 ) .
In a stratified medium composed of a finite number of homogeneous layers overlaying
an homogeneous half-space it may be of interest to evaluate how Rayleigh phase velocity
VR ( ) is affected by a small change in body wave velocities VP and VS of one layer. This
quantity can be computed by integrating Eq. 3.44 over the layer thickness. Thus, for a
layered medium composed by N L layers L j ( j = 1, N L ) the results is:

j
VR
VS

= 2
VS L j 2k UI1 y j1

yj

dr1
dr
kr2 4 kr1 2 dy
dy
dy

(3.45)

VR

VP
dr2
kr1 + dy

= 2
dy
VP L j 2k UI1 y j1

3.5.2

Effective Rayleigh Phase Velocity Partial Derivatives

In Section 3.5.1 closed-form expressions for the partial derivatives of Rayleigh phase
velocity with respect to the medium parameters have been obtained. These partial
derivatives are modal quantities, in the sense that they refer to a specific mode of
propagation of Rayleigh waves. However, it was shown in Section 3.3 that in vertically
heterogeneous media excited by harmonic sources, Rayleigh waves propagate in wave trains
that result from the superposition of different modes of propagation. In these
circumstances, the relevant kinematical quantity that describes the velocity of propagation of
the wave train is the effective Rayleigh phase velocity.
In this section, closed-form solutions for the partial derivatives of the effective Rayleigh
phase velocity V$ (r , y , ) with respect to the body wave velocities of the medium V and
P

VS will be obtained. From Eq. 3.23 the effective Rayleigh phase velocity may be considered

[(

) (

a function of the following variables V$ = V$ r1 y , k j , , r2 y , k j , , Vj , U j , k j , where


= r , y and j = 1, M . The explicit dependence of V$ on the independent variables r and
has been omitted because it is irrelevant in the next developments. It is convenient to rewrite Eq. 3.23 as follows:

84

Rayleigh Waves in Vertically Heterogeneous Media

[ (

( )

)]

cos r 1 V 1 V
i
j
ij

Vi Vj
i =1 j= 1

$
V ( r , y , ) =
M M
rs (1 Vr + 1 Vs ) cos r (1 Vr 1 Vs )

Vr Vs
r = 1 s =1

( )

(3.46)

where:

( )
r

( )
y

( )

ij

( )

ij

(3.47a)

(3.47b)

r ( y , k )r y , k r ( y , k )r y , k
1
i 1
j 2
S
i 2
S
j
=

( U i I i ) U jI j

( )

r (y , k )r y , k r ( y , k )r y , k
2
i 2
j 2
S
i 2
S
j
=

(U i I i ) U j I j

( )

Hence from Eq. 3.46:

[(

) (

V$ = V$ r1 y , k j , , r2 y , k j , , Vj , U j

(3.48)

since Vj and k j = Vj are not independent, and I j is given by Eq. 3.22.


Mimicking the procedure used in Section 3.5.1, the problem of determining the partial
derivatives of V$ with respect to the medium parameters is essentially reduced to that of
computing V$ for small variations of VS and VP . From Eq. 3.48, it is apparent that the

latter task requires the computation of quantities such as (r1 ) j , (r2 ) j , Vj , and U j .
However from Rayleigh principle, first-order perturbations in the wavenumber k j will induce

variations in the corresponding eigenfunctions (r1 ) j and (r2 ) j that are of second order.

Hence, in computing the first variation of V$ , the terms (r1 ) j and (r2 ) j may be

neglected, and from Eq. 3.48:

Rayleigh Waves in Vertically Heterogeneous Media

$
V$

$
V =
Vj +
U j
U j
Vj

85

(3.49)

where the summation convention is implied over the index j. The first variation of Vj is
given by Eq. 3.42, and U j can be computed from Eq. 3.39 resulting in:

(V I3 2I2 )
I
I
I
U j = 2 2 V + 2 3 +
I
V I1 2I1
(2V I12 ) 1 j
V I1

(3.50)

In Appendix B it is shown that the first variations of Vj and U j given by Eq. 3.41
and Eq. 3.50, respectively, can be written as follows:

Vj = PjVS + Q j VP dy
0

(3.51)

U j = j VS + jVP dy
0

(3.52)

where:

Pj ( y ,) =

2 ( k 2 UI1 ) j

Q j ( y ,) =

and

VS

VP

dr1
dr
kr2 4 kr1 2
dy
dy

2 ( k 2 UI1 ) j

(3.53a)

dr
kr1 + 2
dy j

(3.53b)

86

Rayleigh Waves in Vertically Heterogeneous Media

j ( y,) =

2 (I1 ) j

2
dr1
dr2
2

VS kr2 r2
2r1 ( k I2 ) j Pj
dy
dy j

(3.54a)

2
dr2
2
j ( y ,) = 2
VP kr1 r1 ( k I2 ) j Q j
dy j

(I1 ) j
1

(3.54b)

In view of Eqs. 3.51 and 3.52, Eq. 3.49 can be written as follows:

$
$
$

V
V

V$
V

V =
Pj +
Qj +
j VS +
j VP dy
V
U j
U j
Vj

0
j

(3.55)

which, by analogy to Eq. 3.42 suggests the following definition:

$
V =
0

V$
VS dy +

VS ,VP

V$

VP dy

VP ,VS

(3.56)

where:

$
$
V
V$
V
=
Pj +

U j j
VS ,VP Vj
(3.57)

$
V$
V
V$

=
Qj +
j

P ,VS
j
j

Equation 3.57 provides expressions for computing the partial derivatives of the effective
Rayleigh phase velocity V$ (r , y , ) with respect to the medium parameters V and V . To
P

Rayleigh Waves in Vertically Heterogeneous Media

87

complete the formulation requires explicit expressions for the partial derivatives
V$
U j

V$
Vj

and

. Appendix B shows the details of this computation. The final result is:

M M
$
$
V
V

~
~
=
r , y , ) = Pi Pj + T i + T j

(
ij
ij
ij
ij

i =1 j=1
VS ,VP VS

( )

( )

( )

( )

M M
$
$
V
V

~
~
=
r , y , ) = Q i Q j + T i + T j

(
ij
ij
ij
ij

i =1 j=1
VP ,V VP
S

( )

( )

( )

( )

(3.58)

( ) , (~ ) and (T ) , (~T ) are complicated functions of the modal


parameters V , U , r ( y , k , ) and r ( y , k , ) , the frequency of excitation , and the
The terms
j

ij

ij

ij

ij

distance from the source r. Because their expressions are rather lengthy, they have been
reported in Appendix B
The distinctive feature of the effective partial derivatives that distinguishes them from
the modal partial derivatives is their dependence on r. The local properties of these
quantities have been inherited from the effective Rayleigh phase velocity during the process
of differentiating Eq.3.46.
3.5.3

Attenuation of Rayleigh Waves in Weakly Dissipative Media

In the previous sections it was shown how variational methods can be used to obtain
useful expressions regarding Rayleigh wave propagation in elastic media. In particular, these
methods were used to determine closed-form solutions of the partial derivatives of modal
and effective Rayleigh phase velocity with respect to the medium parameters. Another
useful result was an alternate expression for computing Rayleigh group velocity without the
use of numerical differentiation.
Variational methods can also be used to obtain important results for Rayleigh wave
propagation in weakly dissipative media. In Chapter 2 weakly dissipative media were defined
as those media satisfying Eq. 2.31. It was also shown that the mechanical properties
governing the behavior of such media are defined by Eqs. 2.34 and 2.35. In physical terms
these are the phase velocity V ( ) and attenuation coefficient ( ) of compression and
shear waves ( = P , S) . In this section variational techniques are employed to obtain phase

88

Rayleigh Waves in Vertically Heterogeneous Media

velocity VR ( ) and attenuation coefficient R ( ) of Rayleigh waves propagating in weakly


dissipative media. Equation 3.42 also implies the following result:

V
VR = R
VS dy +
0 VS ,VP

VR

V
dy

P
VP ,VS

(3.59)

In weakly dissipative media the material damping ratio D is a small quantity, and hence
it is reasonable to assume that in Eq. 2.34 the second order terms in D are in most cases

negligible. With this assumption Eq. 2.34 simplifies to V* ( ) = V ( ) 1 iD

with

V ( ) given by Eq. 3.35. If Ve denotes the phase velocity of the -wave in a linear elastic
medium, the existence of material damping may be thought (Anderson and Archambeau,
1964; Anderson et al., 1965) to introduce a small change in Ve given by:

[(

V * = V V e iV D

(3.60)

Then, substituting Eq. 3.60 in Eq. 3.59 for VS and VP :

V
VR
e

+
VR* = R
V
V
iV
D
dy
VP VPe ) iVP D P dy
(
(

S
S)
S S

VS ,V
VP ,V
0
0
P
S

*
where VR =

[( V

(3.61)

VRe ) iVR D R . Taking the real and the imaginary parts of Eq. 3.61:

V
VSe
VR VPe
R
VR ( ) = V + VS

1
dy + VP
1
dy
VP ,VS VP
0 VS ,VP VS
0
e
R

VR
VR
1
dy
D R () =
VS D S
dy
+
V
D

P P VP
VR ( ) 0
VS ,VP

0
, VS

(3.62)

Rayleigh Waves in Vertically Heterogeneous Media

From Eq. 2.35, 1

V e 2 D
ln
=
V ( )
ref

89

e
where V ref = V since ref is the

reference frequency for material dispersion. Also, R ( ) =

D from Eq. 2.32. With


VR ( ) R

these results Eq. 3.62 can be rewritten as follows:

2
VR
VR
VS D S
+
VR ( ) = V + ln
dy
V
D
dy

0 P P VP
VS ,VP

ref 0
, VS
e
R

VR
VR

R ( ) =
dy + VP D P

dy
2 VS D S
VS ,VP
VP ,VS
VR ( ) 0
0

(3.63)

Equation 3.63 is an important result because it shows that Rayleigh phase velocity
VR ( ) and attenuation coefficient R ( ) in vertically heterogeneous, weakly dissipative
media can be computed from the solution of the elastic Rayleigh eigenvalue problem. In
fact Eq. 3.63 forms the basis of an algorithm for the solution of the uncoupled inverse
problem of Rayleigh waves. The procedure, which will be described in detail in Chapter 4,
involves three major steps. The first step is the determination of the experimental
dispersion and attenuation curves, namely VR ( ) and R ( ) , from surface wave
measurements. In the second step the experimental dispersion curve VR ( ) is inverted to
obtain the elastic shear wave velocity profile VS ( y ) . The third and final step involves the
use of Eq. 3.63 as the basis of the inversion of the experimental attenuation curve R ( )
to obtain the material damping ratio profile D S ( y) .

It should be noted that the inversion for VS is non-linear, but the inversion for D S ( y)
is linear. Most of the procedures currently used by seismologists to study Rayleigh wave
attenuation are based on Eq. 3.63 and the assumption of weak dissipation (Lee and
Solomon, 1979; Aki and Richards, 1980; Keilis-Borok, 1989; Herrmann, 1994). The next
section will illustrate a new technique that can be used to simultaneously invert both the
dispersion and attenuation curves. The technique is quite general and can also be applied to
strongly dissipative media.
Although Eq. 3.63 has been derived with reference to individual Rayleigh wave modes, it
is also valid for the case of effective Rayleigh wave propagation with small changes of
notation. The primary modification is the replacement of the modal partial derivatives (Eq.
3.44) with the effective partial derivatives (Eq. 3.58). Then the modal phase velocities and
attenuation coefficients VR ( ) and R ( ) must be replaced with the corresponding

90

Rayleigh Waves in Vertically Heterogeneous Media

effective quantities V$ (r , ) and $ (r , ) . However, for the latter operation to be

legitimate, Eq. 2.32b, namely R ( ) = k R ( ) D R ( ) , must be assumed to be true for the


effective (instead of modal) quantities.
3.6 Rayleigh Eigenvalue Problem in Viscoelastic Media
The approach used to define the Rayleigh eigenvalue problem in elastic media was based
on the application of Lagranges equations to a vertically heterogeneous elastic medium. A
solution of the resulting Naviers equations of motion was then sought in a form of a
harmonic displacement field satisfying the boundary conditions for surface waves. The
generalization of this procedure to viscoelastic media requires the use of certain variational
theorems of linear viscoelasticity (Christensen, 1971). The equations of motion and the
associated boundary conditions are then established from the stationary condition of an
energy functional.
However, integral transform methods provide a more straightforward approach to the
formulation of the Rayleigh eigenproblem in viscoelastic media. For boundary value
problems with time-invariant boundary conditions, the use of integral transform methods
leads naturally to the application of the elastic-viscoelastic correspondence principle.
According to this principle, elastic solutions to steady state harmonic problems can be
converted into viscoelastic solutions for identical boundary conditions by simply replacing
the elastic moduli with the corresponding frequency dependent complex moduli (Read,
1950; Christensen, 1971). Application of the correspondence principle to the elastic
eigenproblem leads to a complex eigenproblem for viscoelastic media. Formally, Eq. 3.10
and the associated boundary conditions (Eq. 3.11) are still adequate to describe this complex
eigenproblem if the vector f ( y ) and the matrix A ( y) are complex-valued arrays. Most of
the features described for the elastic eigenproblem are carried over to the complex
eigenproblem with the important difference that non-trivial solutions of the latter are
complex-valued eigenvalues (i.e the wavenumbers) and eigenfunctions. It should be
remarked, however, that certain properties of the real eigenproblem require further
consideration before they can be generalized to the complex eigenproblem. One example is
the orthogonality of the eigenfunctions.

The techniques available for the solution of the complex eigenproblem are in essence
the same as those used for the elastic eigenproblem. The main difference is that the use of
complex arithmetic can no longer be avoided in solving the viscoelastic eigenproblem, and
algorithms such as root finding techniques must be properly generalized to remain
applicable for complex values of the arguments. This generalization is not always trivial as
shown in the next section.
Before concluding this section it is worth noting that the viscoelastic eigenproblem
includes an interesting degenerate case. If the complex Lams parameters are such that the

Rayleigh Waves in Vertically Heterogeneous Media

91

viscoelastic Poissons ratio is a frequency-independent real number, it can be shown that the
roots of the Rayleigh dispersion equation as well as the corresponding eigenfunctions are
real (Christensen, 1971). Therefore, in this special circumstance, the solution of the complex
eigenproblem could be obtained using the same procedures used for the elastic
eigenproblem. The wavenumbers for the viscoelastic eigenproblem will still be complex,
however.
3.6.1

A Solution Technique

A new technique has been developed for the solution of the complex eigenvalue
problem. This technique is based on the generalization to viscoelastic media of the method
developed by Chen (1993) and Hisada (1995) for finding the normal modes in multi-layered
elastic half-spaces. In their algorithm the authors solve the elastic eigenproblem using the
method of reflection and transmission coefficients. Figure 3.6 shows the parameters
characterizing the properties of viscoelastic multi-layered media.

Rayleigh Wave

r
Layer 1

h 1 , 1 , V P 1 , V S1 , D P 1 , D S1

Layer 2

h 2 , 2 , V P 2 , V S2 , D P2 , D S2

M
Layer nl

Figure 3.6

h n l , n l , V Pn l , V Sn l , D Pn l , D Sn l

Rayleigh Waves in Viscoelastic Multi-Layered Media

As mentioned in Section 3.2.1, most of the computational effort required for the
solution of the elastic eigenvalue problem are spent in constructing the Rayleigh secular
function and in finding its roots. This is also true for the viscoelastic eigenproblem with new
difficulties arising in connection with the root computation of the Rayleigh secular function.
The latter is now a complex-valued function F R () of the complex-valued Rayleigh phase

92

Rayleigh Waves in Vertically Heterogeneous Media

velocity VR* defined by Eq. 2.34: symbolically F R ( VR* ): C C (for simplicity, the use of an
asterisk to denote a complex quantity will hereafter be limited to the essential cases).
In general, the problem of computing the roots of a complex-valued function f ( ) of a
complex variable z, namely f ( z): C C , is not simple, particularly if the function is highly
non-linear and is known only numerically (Henrici, 1974). No general methods are currently
available, and in most cases the adopted strategy consists in breaking up the complex
structure of the function and transforming it into an equivalent pair of non-linear equations
in two real variables. The root-finding problem is then addressed by conventional methods.
When this approach was used in this study to find the roots of the real and imaginary parts
of the Rayleigh secular function, it was found to have limited success, and only for very
small values of material damping ratio.
Most of the difficulties associated with computing the roots of functions f ( z): C C
are overcome if the function f ( z) satisfies the condition of analyticity within an open set
D C (see Hille, 1973 for a precise definition of the necessary and sufficient conditions
required for a function f ( z): C C to be analytic or holomorphic in D C ). In this case the
zeros of the function f ( z) may be determined using a completely new class of algorithms
which are developed by taking full advantage of the theory of analytic functions. Strictly
speaking, the Rayleigh secular functionF R ( VR* ) is not analytic with respect to the complex

variable Rayleigh phase velocity VR* , because it is not a single-valued function of VR* .
However, this fundamental property can be restored (within a set D : VR* D C ) during
the process of constructing F R ( VR* ) by choosing the appropriate branches of F R ( VR* ) on
the Riemann surfaces. The practical implementation of this process requires that all of the

branch-cuts and branch-points ofF R ( VR* ) in D be identified first (Bth, 1968; Schwab and
Knopoff, 1972).
In this study a new technique for computing the zeros of F R ( VR* ) was developed. The
technique finds the roots of the Rayleigh secular function without breaking up its complex
structure. Conversely, it takes advantage of the intimate connection existing between the
real and the imaginary part of F R ( VR* ) , as a consequence of its analyticity with respect to
VR* in the domain of interest, with the exception of at most a finite number of isolated
pole-type singularities. In this case the term meromorphic function would be more appropriate
than holomorphic function to designate F R ( VR* ) . The proposed method is based on the
theory of analytic functions, particularly the well-known Cauchys residue theorem:

Rayleigh Waves in Vertically Heterogeneous Media

1
f( z )dz =
2 i

Re s(z )

93

(3.64)

j= 1

where the integral sign denotes integration along a positively oriented closed contour ,
f ( z) is an analytic function inside and on except at the points z j ( j = 1, M) where it may

have isolated singularities; z C is a point of the complex plane z = ( x + iy ) , and the

()

symbol Res z j denotes the residue of the function f ( z) at the point z j . Finally, M is the
number of isolated singularities of f ( z) located inside . Equation 3.64 forms the basis of
the algorithm proposed by Abd-Elall et al. (1970) for computing the roots of f ( z) . In fact
Cauchys residue theorem may also be written in the form:

GN =

zN
1
dz =

2 i f( z )

z
j= 1

j j

(3.65)

where j are the residues of 1 f( z) at the points z j that are the zeros of f ( z) , m is the

number of zeros of f ( z) which are located inside . By evaluating the contour integral
defined by Eq. 3.65 for different values of N ( N = 0, 2 m 1) , a sequence of complex
numbers G N
G
determine the coefficients of the complex polynomial Pm ( z) :
Pm ( z) = c 0 + c1 z + c 2 z 2 +....+ c m 1 z m 1 + z m

(3.66)

by solving the linear system of equations that can be constructed from the modified
Newton identities (Abd-Elall et al., 1970):
m 1

G
j= 0

r+j

c j + G r +m = 0

r = 0, m 1

(3.67)

The zeros of the polynomial (Eq. 3.66) coincide with the zeros of f ( z) inside . It
should be remarked that solution of the system of equations (Eq. 3.67) does not require
knowledge of the residues j .
If f ( z) is identified with F R ( VR* ) , and the contour is the boundary of a region D

where the roots of F R ( VR* ) are located, then the roots of Pm ( z) can be interpreted as the

94

Rayleigh Waves in Vertically Heterogeneous Media

modal (complex) Rayleigh phase velocities associated with the solution of the complex
eigenproblem. A fundamental step required for the implementation of the algorithm is the
computation of the complex numbers G N for N = 0, 2 m 1 . This task is accomplished by
evaluating the contour integral of Eq. 3.65 numerically.
A crucial step of this calculation is the definition of the region D , delimited by the

boundary , where the roots of F R ( VR* ) are located. To define D , one must determine
lower and upper bounds for the real and imaginary parts of VR* . In the plane

w R : VR , D R lower and upper bounds for VR may be easily established from the roots of
the Rayleigh dispersion equation in homogeneous media, using min( VS ) and max ( VS ) ,
respectively. Bounds for
DR
are found from the observation that
0 D R max (D R ) < max ( D S ) . However, the plane of analyticity of the Rayleigh secular

function is not the plane w R : VR , D R

but rather the plane z R : x R , y R , whose

relations with w R : VR , D R are given by Eq. 2.34, namely:


VR

=
=
x
x
V
,
D
(
)
R
R
R
R

(1 + D R 2 )

zR:
VR D R

=
=

y
y
V
D
,
(
)
R
R
R
R

(1 + D R 2 )

(3.68)

such that VR* = z R = ( x R + iy R ) . In Eq. 3.65, the numerical evaluation of the integral
involving F R ( VR* ) along the contour must be performed in the zR-plane.

Equations 3.68 give the relationships between a point w R : ( VR , D R ) of the wR-plane and

the corresponding point z R : ( x R , y R ) of the zR-plane. Thus, it is possible to map a region


C of the wR-plane into another region D of the zR-plane. Figures 3.7(a) and (b) show the
two regions C and D where the roots of F R ( VR* ) are located. It should be noted from
the figures the opposite orientation of the boundaries and of the two regions C and D
of the complex plane.
In Eq. 3.65 the numerical evaluation of the contour integral may be simplified by
introducing an admissible parametrization of the contour so that the numbers G N may
be computed as follows:

Rayleigh Waves in Vertically Heterogeneous Media

95

10
8

wR-plane

DR (%)

roots
D

0
-2
0

100

200

300

400

500

600

VR (m/s)

Figure 3.7(a)

Roots of Rayleigh Secular Function in the Region C of the wR-plane


5

YR (m/s)

roots

-5

-10

-15

-20
0

zR-plane

100

200

300

400

500

600

XR (m/s)

Figure 3.7(b)

Roots of Rayleigh Secular Function in the Region D of the zR-plane

96

Rayleigh Waves in Vertically Heterogeneous Media

z R ( )
z NR
1
1
GN =

dz R =

z ( ) d
2 i F R ( z R )
2 i F R z R ( ) R

(3.69)

where R is a parameter. A necessary condition for an admissible parametrization z R ( )


of a curve is the existence of the function z R ( ) . This requirement is not satisfied if the
curve is parameterized with the Cartesian representation that follows from Eq. 3.68
because of the existence of a pole-singularity at the points A and B of the zR-plane. Using a
polar rather than a Cartesian representation of overcomes this difficulty. An admissible
parametrization of the contour is therefore given by following equations:
arc AB

arc BC

:
arc CD

arc DA

z() =

VR min VR max

z() = VR max cos2 ( ) i sin(2)


2

0 a tan(DR max )

z() =

1 + D2 R max

[ 1 + iDR max ]

z( ) = VR min cos 2 ( ) + i sin(2)


2

VR max
1 + D2 R max

VR min

(3.70)

1 + D2 R max

a tan(DR max ) 0

where VR min = min( VR ) , VR max = max ( VR ) , and D R max = max ( D R ) . In calculating the
integral of Eq. 3.69 care must be used because the two curves (wR-plane) and (zR-plane)
are oriented in opposite directions.
A substantial improvement in the accuracy of the above numerical integration can be
achieved by rescaling the contour of integration (zR-plane) via a conformal linear
transformation of the type z R = ( R R + R ) where R C and R , R R are two
constants (this type of mapping is a special case of a Mbius transformation). Under this
conformal mapping the region D of the zR-plane is mapped into a similar region D' of the
R -plane. Depending on the value adopted for R , the region D' is the magnified or
contracted version of D . The factor R is responsible for shifting the region D from the
origin. After some experimentation it was found that the optimum values for R and R
are R = max (VR ) and R = 0 .

The method of substitution in C , applied to the contour integral of Eq. 3.69 yields:

Rayleigh Waves in Vertically Heterogeneous Media

97

[zR ()] z ()d = R [ R R (') + R ] (')d'


1
GN =

R
R
2 i F R [z R ()]
2i ' F R [ R R ( ') + R ]
N

(3.71)

where the contour ' denotes the boundary of the region D' in the R -plane. The
numerical integration in Eq. 3.71 was performed using a classical fifty-point Gauss-Legendre
quadrature formula (Press et al., 1992).
Once the numbers G N ( N = 0, 2 m 1 ) have been calculated, the remaining steps
required for the complete implementation of the algorithm are 1) the solution of the linear
systems of equations (Eq. 3.67) for the coefficients c j ( j = 0 , m 1) , and 2) the computation

of the roots of the complex polynomial Pm ( z) of Eq. 3.66. The task of solving the linear
system of equations is accomplished with the complex version of the LU decomposition
algorithm, whereas the roots of Pm ( z) are computed using La Guerres method
supplemented with the appropriate deflation and polishing techniques (Press et al., 1992).
At each frequency , the determining the number of zeros of F R ( VR* ) is the first step
in finding the roots of the Rayleigh secular function. A possible strategy for computing m is
based on the observation that all the matrices G j with j > m are singular (Abd-Elall et al.,
1970). Hence, the value of m can be found by a procedure that evaluates the rank of
successive matrices G j until a value of j is found for which G j is singular. In this study the
test for singularity of the matrices G j was performed using their condition number R cond that
was calculated using the singular value decomposition of G j . A matrix G j is considered
singular if its condition number is so small that the logical expression 1.0 + R cond = 1.0 is
true to machine precision.
If the number of roots of F R ( VR* ) which is equivalent to the number of Rayleigh
modes M (M = m) associated with a frequency is large, say mmax greater than 15 or 20,
the computation of the roots of the high-degree polynomial Pm ( z) is an ill-conditioned
problem. As a result, it may be difficult to compute the roots ( VR* ) j ( j = 1, m ) of the

Rayleigh secular function with a high degree of accuracy. A strategy indicated by Delves and
Lyness (1967) to overcome this problem is to subdivide the region D' into smaller

subregions where the number of roots of F R ( VR* ) in each subregion is less than mmax . The
problem of finding the roots of the Rayleigh secular function in D' is then reduced to that

of computing the roots of F R ( VR* ) in each subregion. The calculation of the zeros of

98

Rayleigh Waves in Vertically Heterogeneous Media

Pm ( z) completes the problem of finding the eigenvalues

(k )
R

= ( VR* ) j ( j = 1, M)

associated with the solution of the complex eigenproblem.


Once the roots of F R ( VR* ) have been computed for a given frequency , the
remaining task is to determine the corresponding (complex) eigenfunctions ri* y , ( k R ) j ,

( i = 1, 4 ) . The technique used in this study to compute ri* y , ( k R ) j , is the same

technique used to solve the elastic eigenproblem, and is based on the algorithm developed
by Chen (1993) and Hisada (1995). The only difference in computing the elastic and the
viscoelastic eigenfunctions is that, in the latter, all the operations must be performed using
complex arithmetic. However, the fact that the Rayleigh eigenfunctions are complex-valued
for viscoelastic media implies that the phase difference between the horizontal and the
vertical components of the displacement field is no longer equal to 2 as in the elastic
case. As a result, the principal axes of the ellipse describing the trajectory of the Rayleigh
wave particle motion are rotated forward or backward with respect to the free surface of the
half-space (Bth, 1968).
Determination of the wavenumbers

(k )
R

= ( VR* ) j

and of the associated

eigenfunctions ri* y , ( k R ) j , completely solves the complex eigenproblem and thus the

homogeneous boundary value problem of Rayleigh waves in viscoelastic, vertically


heterogeneous media. The corresponding dispersion and attenuation curves can be easily
obtained from the inversion of Eq. 3.68 to give:

(x R 2 + y R 2 )
VR = VR ( x R , y R ) =
xR

wR:

yR
D R = D R ( x R , y R ) = x
R

(3.72)

where VR = VR (k R ) j , , and D R = D R ( k R ) j , with j = 1, M . To properly account

for causal material dispersion, the Rayleigh secular function F R ( VR* ) must be constructed
using body wave velocities VP = VP ( y , ) and VS = VS ( y , ) calculated with Eq. 2.35.

Rayleigh Waves in Vertically Heterogeneous Media

99

3.7 Effective Phase Velocity and Greens Function in Viscoelastic Media


The purpose of this section is to generalize the results obtained in Sections 3.3 and 3.4
for elastic materials to viscoelastic media. The main tool used to implement this
generalization will again be the elastic-viscoelastic correspondence principle. The

displacement field u (r , y , ) ( = r , y ) induced by a harmonic point source Fy e it in a


vertically heterogeneous elastic half-space is given by either Eq. 3.28 or Eq. 3.12. The latter
representation follows directly as solution of the inhomogeneous boundary value problem
of Rayleigh waves in elastic, vertically heterogeneous media (Aki and Richards, 1980).
An efficient approach for determining the solution of the corresponding problem in
viscoelastic media involves the use of integral transform methods. The governing equations
obtained with this procedure will be formally identical to those of the elastic problem,
except that the transformed viscoelastic field variables replace the elastic field variables, and
the viscoelastic complex moduli replace the elastic moduli (if using the Fourier transform).
This association between elastic and viscoelastic solutions is the essence of the
correspondence principle that, if applied to Eq. 3.12, yields:

u (r , y , ) =

[ A ( r , y , )] e (

and k *j ( ) are the (complex) Rayleigh displacement amplitude and

where A * (r , y , )

j= 1

i t k *j r +

(3.73)

wavenumber, respectively, associated with the jth mode of propagation ( j = 1, M) . Equation


3.73 allows the displacement field induced by an harmonic point source Fy e it in a
vertically heterogeneous, viscoelastic half-space to be determined. The modal amplitude

[A (r , y , )]
*

can be determined from Eq. 3.21 if all of the modal quantities are replaced

with those obtained from the solution of the complex eigenproblem, namely

{k , r (y, k ,)} (i = 1, 4) . Equation 3.73 also shows that the viscoelastic solution can be
*
j

*
i

*
j

obtained via an extension of the elastic solution (3.12) to complex values of

{k , r (y , k ,)} . It should be remarked that this extension is restricted only to those


j

subsets D1 D2 ... Dn C where Eq.3.73 is a well-defined, single-valued and


continuous function of the independent complex arguments.
As already mentioned, in elastic media Eqs. 3.12 and 3.28 provide equivalent

representations for u (r , y , ) . Therefore, following the procedure that led to Eq. 3.73, the

100

Rayleigh Waves in Vertically Heterogeneous Media

viscoelastic Greens function may be obtained from Eq. 3.24 by extending the validity of the
elastic Greens function to complex valued amplitude and phase to yield:
i t * ( r ,y , ) ]
u$ (r , y , ) = U$* (r , y , ) e [

(3.74)

where the terms U$* , * C can be computed using Eqs. 2.25 and 2.26 with complexvalued modal quantities. The restrictions about the domains of validity of Eq. 3.73 apply
equally well to Eq. 3.74. However, the presence in Eqs. 2.25 and 2.26 of rather complicated
multi-valued functions of several complex variables can make the task of identifying branch
cuts, branch points and singularities associated with these expressions very difficult (Krantz,
iarg (U$* )
Eq. 3.74 may be more conveniently re-written as:
1982),. By setting U$ * = U$ * e

i t *v ( r , y , ) ]
u$ (r , y , ) =U$v (r , y , ) e [

(3.75)

where the terms U$v R , v* C are defined respectively by U$v (r , y , ) = U$* and

( )]

*v (r , y , ) = * arg U$* .
Finally, from Eq. 3.75 the displacement field u (r , y , ) induced by a harmonic source

Fy e it located at r = 0, y = y S
can be written in a form:

in a linear viscoelastic vertically heterogeneous medium

u (r , y , ) = Fy Gv (r , y , ) e [

i t *v ( r , y , )

(3.76)

where Gv (r , y , ) =U$v (r , y , ) is the Rayleigh geometric spreading function in viscoelastic


media, which is in general different from G (r , y , ) .

From Eq. 3.76, the expression t *v (r , y , ) = cons tan t represents the equation of
a (complex) wave front, which is characterized by an effective complex phase velocity given
by:

Rayleigh Waves in Vertically Heterogeneous Media


V$ * (r , y , ) =

where *v (r , y , )
by k$ * (r , y , ) .

*v (r , y , )

(3.77)
,r

can be interpreted as an effective complex wavenumber and denoted

,r

To

obtain quantities of physical significance

k$ * (r , y , ) =
+ i $ (r , y , ) so that:
V$ (r , y , )

V$ (r , y , ) =

101

[( )

( )]

* arg U$*

[ ( )]

$ (r , y , ) = *

it

is

convenient

to

set

(3.78a)
,r

,r

(3.78b)

V$ (r , y , ) and $ (r , y , ) are the effective phase velocity and the effective attenuation
coefficient, respectively, and characterize the harmonic propagation of multi-mode Rayleigh
waves in viscoelastic, vertically layered media. At a fixed y = y c , the two-dimensional plots
V$ (r , y c , ) and $ (r , y , ) are defined as the dispersion and the attenuation surface,

respectively. It should be remarked that Eq. 3.78 could have also been obtained from the
extension of the elastic solution (i.e. Eq. 3.23) for complex values of the arguments.
3.8 Modal and Effective Partial Derivatives in Viscoelastic Media
The results presented in Section 3.5, particularly the closed-form partial derivatives of
Rayleigh phase velocity with respect to medium parameters were obtained using the
Hamilton variational principle. The applicability of this fundamental principle of dynamics is
restricted only to monogenic mechanical systems, i.e., to those systems for which the internal
and external forces are derivable from scalar potentials which are functions of particle
coordinates, particle velocity and time (Lanczos, 1970; Goldstein, 1980). Elastic bodies
subjected to specified distributions of body forces and surface tractions fulfill the conditions
set forth for monogenic systems. In fact, they satisfy the even more restrictive conditions of
conservative systems, for which the potential functions (in this case the strain energy) do not
have an explicit dependence on particle velocity and time. For monogenic continuous

102

Rayleigh Waves in Vertically Heterogeneous Media

systems, the Lagrangian density is simply given by the difference between kinetic energy
density and potential energy density.
Viscoelastic bodies, on the other hand, belong to the category of dissipative systems
that are not monogenic because dissipative forces do not admit a representation in terms of
potential functions. As a result, Hamiltons variational principle is not applicable in
viscoelastic bodies, at least in its classical form. These considerations seem to suggest that
the important results obtained in Section 3.5 and derived from the application of
Hamiltons principle do not hold in viscoelastic media.
Two approaches may be used to show that, fortunately, this is not the case. The first
approach is based on introducing a type of formalism that allows treating non-conservative
systems as if they were conservative (Ben-Menahem and Singh, 1981). The basic idea is to
consider a mechanical system made up from the combination of two systems: the real
system A , which stores and dissipates strain energy, and a mirror system B producing
energy in equal amount to that dissipated by A (i.e. the system B acts as a storage for the
energy dissipated by A ). For the joined system A U B the total energy is conserved, and
hence Hamiltons principle can be applied. The results of this approach show that the
variational principles of Rayleigh and Love waves along with their implications are also valid
for viscoelastic media, and thus Eqs.3.44 and 3.58 remain valid for complex values of the
parameters. However, the analysis of Ben-Menahem and Singh (1981) indicates that the
results are correct only to a first order approximation in D ( = P , S) , and hence they are
rigorously valid only in weakly dissipative media.
The second approach that may be used to show that the Rayleigh variational principle,
can be extended to viscoelastic media is based on the application of certain variational
theorems of linear viscoelasticity (Gurtin, 1963; Christensen, 1971). As mentioned in
Section 3.6, these theorems can also be used as an alternative to integral transform methods
to obtain the equations of motion and the associated boundary conditions of various
boundary value problems in linear viscoelasticity. Most of the viscoelastic variational
theorems are natural extensions of the results obtained in linear elasticity. They provide the
rigorous procedure for extending the variational theorems of linear elasticity to
viscoelasticity, in particular the Rayleigh variational principle, which is formally expressed by
Eqs. 3.34 and 3.35.
However, as remarked several times in this chapter, integral transform methods are an
ideal tool for solving boundary value problems in linear viscoelasticity because of the
advantages offered by the application of the elastic-viscoelastic correspondence principle.
Among them is the possibility of reinterpreting the variational theorems of linear elasticity
as integral transformed viscoelastic variational theorems (Christensen, 1971). For steady
state harmonic problems, this reinterpretation becomes a trivial exercise simply involving
the extension of the elastic solution to complex values of the field variables. Once the
viscoelastic version of the Rayleigh variational principle has been established, the results

Rayleigh Waves in Vertically Heterogeneous Media

103

obtained in Sections 3.5.1 and 3.5.2, particularly Eqs. 3.34 and 3.35, can be extended to
linear viscoelasticity on a firm theoretical basis.
The approach based integral transform methods does not have any physical
interpretation and is solely based on the formal analogies between the field equations of
linear elasticity and the integral transformed field equations of linear viscoelasticity. On the
other hand, the approach based viscoelastic variational theorems has both a solid theoretical
basis and a physical, energy-based interpretation. However, the thermodynamical legitimacy
of the procedure based on the application of Hamiltons principle to the combined realmirror system (Ben-Menahem and Singh, 1981) is questionable. Furthermore, this approach
lead to results that are correct only to first order.

104

Rayleigh Waves in Vertically Heterogeneous Media

4 SOLUTION OF THE RAYLEIGH INVERSE PROBLEM

4.1 Introduction

Given the set of medium parameters ( y ), VP* ( y ), VS* ( y) defining the material
properties of a site, the problem of determining the dispersion and attenuation curves
VR ( ) and R ( ) associated with that site is often referred to as the Rayleigh direct or
forward problem. Conversely, if VR ( ) and R ( ) are known, then the problem of

determining the unknown medium parameters ( y ), VP* ( y ), VS* ( y) defines the Rayleigh
backward or inverse problem.

In more general terms, direct problems are concerned with determining the effects
induced on a physical system by certain causes, whereas in inverse problems the roles of
causes and effects are reversed, and the objective is to determine the causes that generate
the observed effects (Engl, 1993). Following this definition, for a vertically heterogeneous
viscoelastic medium excited by a harmonic source, several types of Rayleigh direct/inverse
problems may be considered. They differ from each other in the function (i.e., the effect)
that is chosen to represent the medium response. Dispersion and attenuation functions are
one possible type of response function but other choices are possible as well. In the frequency
domain other response functions include the displacement amplitude, the displacement
phase, or the displacement spectra. In the time domain and for transient sources, a valuable
response function is either a short or long period seismogram. Although in geotechnical
earthquake engineering the most common response functions are experimentally
determined dispersion and attenuation curves (Stokoe et al., 1989; Tokimatsu, 1995; Rix et
al., 1998a), other interpretations of the test using alternate response functions are also
possible as shown later in this chapter.
Inverse problems concerned with the determination of a model physical parameters
from the measurements of certain model field variables constitute so-called parameter
identification problems (Engl, 1993). In the solution of parameter identification problems, the
ability to successfully invert the measured field variables to obtain reliable estimates of the
model parameters depends to a significant degree on the choice of the response function.
Factors governing the selection of appropriate response functions include the ability to
experimentally measure the field variables, the ability to solve the corresponding inverse
problem (e.g., the procedure used to compute the partial derivatives of the response
function with respect to the model parameters), and the information content associated
with the selected response function.

105

106

Solution of the Rayleigh Inverse Problem

4.2 Ill-Posedness of Inverse Problems


Inverse problems are often inherently ill-posed or unstable, particularly non-linear
problems of parameter identification such as the Rayleigh inverse problem. According to
Hadamards definition of well-posedness, a mathematical problem is said to be well-posed or
stable if it satisfies the following three conditions (Tikhonov and Arsenin, 1977; Engl, 1993):
a. For all admissible data, a solution exists.
b. For all admissible data, the solution is unique.
c.

(4.1)

The solution depends continuously on the data.

To be more precise, the definition (4.1) should also specify the functional space in
which the solution is supposed to exist, and the restrictions that a given set of data must
satisfy to be considered admissible. It should be remarked that the Hadamards postulates of
well posedness apply to both direct and inverse problems. However, it is only in recent
years through studies in non-linear dynamics that the importance of instability has been
recognized in the solution of forward problems where small perturbations in the initial data
produce unpredictable changes in the solution (Parker, 1994).
In inverse problems, conditions b. and c. in (4.1) are often violated. Of particular
relevance in parameter identification problems is the violation of condition b., that is the
existence of more than one solution. For the Rayleigh inverse problem this implies that a
given experimental dispersion curve may correspond to more than one body wave velocity
profile, or analogously, that an experimental attenuation curve may correspond to two or
more different material damping ratio profiles. From a mathematical point of view, nonuniqueness in the solution of an inverse problem is caused by a lack of sufficient
information to constrain the solution. Alternatively, the available information available may
not be independent.
Two strategies can be used to effect uniqueness in the solution of an inverse problem.
The first strategy is to add a priori information about the solution of the problem. For the
Rayleigh inverse problem, this may be information about the body wave velocity and/or
material damping ratio of one or more layers (obtained, for example, from laboratory tests).
Adding constraints to the solution is a second widely used strategy to effect uniqueness
in the solution of an inverse problem. In some cases, the strategy of providing more
information can be regarded as constraining the solution. An obvious example is to require
the body wave velocities and/or material damping ratios to be within a specified range (e.g.,
non-negative). However, there are constraints of a different nature that enforce features of
global behavior such as smoothness and regularity rather than requiring the solution to
assume specific numerical values or bounds.

Solution of the Rayleigh Inverse Problem

107

In discussing strategies to address non-uniqueness problems, it should be noted that the


methods for effecting uniqueness are relatively simple for ideal error-free observations, but
the situation is more complicated for data containing bias and random errors. The next
section will illustrate the most important aspects of the constrained algorithm used as a basis
of the inversion of surface wave experimental data.
Violation of condition c. in (4.1) is also an important concern. Certain inverse problems,
like the solution of the Fredholm integral equation of the first-kind, are very sensitive to
perturbations in the data. Their instability is inherent to the problem and does not depend
on the algorithm used to solve the inverse problem. For linear inverse problems with both
discrete and continuous linear operators, a stability analysis can be carried out by means of
the singular-value expansion method (Menke, 1989; Engl, 1993). The results of this analysis show
that the smallest singular value controls the amplification of the measurement errors. The
rate of decay of the singular values arranged in order of decreasing magnitude is used as a
measure to quantify the degree of instability of a given inverse problem.
For very unstable problems there are mathematical techniques, called regularization
methods, that approximate the ill-posed problem with a parameter-dependent family of
neighboring well-posed problems (Tikhonov and Arsenin, 1977; Engl, 1993). Because some
of these regularization methods admit a variational formulation (e.g. Tikhonov
regularization) where the objective is the minimization of appropriate functionals, they can
also be applied to non-linear inverse problems successfully.
4.3 Coupled Versus Uncoupled Analysis
For the Rayleigh inverse problem, the violation of condition c. in (4.1) is not an issue of

great concern, at least for low to moderate frequencies of excitation (< 50 Hz) . At higher
frequencies an uncoupled analysis, where the dispersion and the attenuation curves are inverted
independently, becomes increasingly sensitive to data perturbations. This problem can be
overcome, or at least mitigated, by using a constrained smoothed inversion algorithm that
has the remarkable property of acting as a regularization method while effecting uniqueness
in the solution. It is interesting to note, however, that a coupled analysis, where the dispersion
and the attenuation curves are inverted simultaneously, is more stable than the uncoupled
analysis, even at high frequencies. The following paragraphs explain this somewhat
unexpected result.
An uncoupled inversion of dispersion and attenuation data requires the solution of two
inverse problems for 2 n L unknown model parameters, for example the shear wave velocities
and the shear damping ratios of a nL-layer soil deposit. The solution of these two inverse
problems is not completely independent because the shear wave velocity profile obtained
from the non-linear inversion of the dispersion curve will subsequently be used in the linear
inversion of the attenuation curve. Therefore, the amplification of the errors resulting from

108

Solution of the Rayleigh Inverse Problem

inversion of the dispersion curve will carry over to the inversion of the attenuation curve, a
process that is characterized by its own degree of ill-posedness. In other words, the
uncoupled inversion suffers a negative synergetic effect resulting from the solution of two
inverse problems where the input data of one problem comes from the solution of the
other.
Conversely, the simultaneous inversion of both the dispersion and attenuation curves
eliminates this negative coupling effect because both sets of experimental data are inverted
simultaneously in a single, complex-valued, inversion. Furthermore, the solution of the
coupled inverse problem takes advantage of an internal constraint that is embedded in the
formalism of the complex inversion. This internal constraint is given by the CauchyRiemann equations that are satisfied by the Rayleigh phase velocity when viewed as an
analytic function of the complex-valued shear wave velocity. The intimate connection
between the real and the imaginary parts of the variables involved in the simultaneous
inversion adds a salutary built-in constraint that makes the coupled inversion a better-posed
problem.
In summary, application of complex variable theory to the simultaneous inversion of
surface wave data is not only an elegant procedure that accounts for the coupling between
elastic moduli and dissipative properties of viscoelastic media (as shown in Chapters 2 and
3), but it also improves the well-posedness of the inverse problem. It should also be
remarked that the simultaneous inversion, in contrast to the uncoupled analysis, is not
restricted by the assumption of weak dissipation.
4.4 Inversion Strategies
In most cases the solution of an inverse problem belonging to the class of parameter
identification problems can be obtained from the solution of an optimization problem which
involves finding the stationary condition of an functional subjected to various constraints
(Parker, 1994). The techniques used to solve non-linear optimization problems can be
broadly divided into global-search procedures and local-search procedures. This distinction
is motivated by the fact that a non-linear optimization problem will have, in general, several
stationary points in the solution space.
Local-search procedures are iterative schemes that, starting from an initial guess of the
solution, generate a sequence of improved approximations converging, under suitable
conditions, toward a stationary point. Most local-search procedures are calculus based
techniques whose strategy consists of locally linearizing a non-linear functional at each
iteration. These techniques require the functional to be sufficiently smooth so that its
Gateaux derivatives with respect to the model parameters exist and are continuous.
Furthermore, even if all the smoothness requirements for the functional are satisfied, the
sequence of approximations of the solution is guaranteed to converge only if the initial
guess is sufficiently close to the solution. However, the most important limitation inherent to

Solution of the Rayleigh Inverse Problem

109

all the local-search procedures is that even when they succeed in finding a stationary point,
there are no simple means to determine whether it is a local or a global stationary point in
the solution space.
This dilemma is addressed by global-search procedures, which are optimization
techniques where the search for a global stationary point is conducted over the entire
solution space. The strategy adopted in a global search method varies according to different
philosophies, some of which include genetic algorithms, fractal inversion, neural network
inversion, enumerative methods, and Monte Carlo simulation. Global-search procedures are
in general more expensive than local-search procedures, both in terms of time and
computer resources. However, they are more robust and reliable compared to the latter.
Figure 4.1 illustrates some of the possible approaches to the Rayleigh inverse problem.
The shaded boxes indicate the options that were considered in this study.

Inverse Problem of Rayleigh Waves


Type of Inversion

Type of Analysis

Global-Search-Methods

Uncoupled Analysis

Local-Search-Methods

Coupled Analysis

Unconstrained
Optimization
Constrained
Optimization

Type of Response Function


Frequency Domain
Dispersion and
Attenuation Functions

Time Domain
Short/Long Period
Seismograms

Displacement Functions
[Complex Spectra]

Occam's Algorithm

Figure 4.1

Algorithms for the Solution of the Rayleigh Inverse Problem

The Rayleigh inverse problem was solved using a local-search procedure, where the
stationary point in the solution space was sought with a constrained optimization technique
known as Occams algorithm. The implementation of the procedure requires the application of
a two-step iterative scheme. During the first step the Rayleigh forward problem is solved for
the current values of the medium parameters. In the second step, the non-linear inverse
problem is linearized in the neighborhood of the current medium parameters so that the
resulting constrained linear inverse problem can be solved. The procedure is repeated for a
sufficient number of iterations until a properly defined convergence criterion is satisfied.
The practical implementation of Occams algorithm relies on the ability to solve two
crucial problems: the Rayleigh forward problem and the computation of the partial

110

Solution of the Rayleigh Inverse Problem

derivatives of the response function with respect to the medium parameters. When the
response functions are the dispersion and attenuation curves, the solution of both problems
has been described in Chapter 3.
4.5 Occams Algorithm
The strategy of solving inverse problems, particularly parameter identification problems,
using constrained optimization algorithms is not new (Lawson and Hanson, 1974) and it is
largely motivated by the need to effect uniqueness in the solution of problems with
uncertain data (i.e., data containing bias and random errors). However, inversions
performed with conventional least-squares techniques where solution is constrained to have
the minimum norm (e.g. the stochastic damped least-squares algorithm) are often
inadequate, and may lead to physically unreasonable profiles of model parameters
(Constable et al., 1987; Menke, 1989). The inadequacy of this class of algorithms can be
attributed to a lack of physical justification for assuming the minimum norm constraint.
A more reasonable approach for constraining the solution of an inverse geophysical
problem is the so-called Occams inversion (Constable et al., 1987; Parker, 1994). The strategy
of this algorithm can be summarized as follows: given a set of experimental data and their
associated uncertainties, find the smoothest profile of model parameters subject to the
constraint of a specified misfit between observed and predicted data. The development of
this class of algorithms was motivated by the following observations. The solution of a
parameter identification problem relies on the ability to synthetically reproduce a set of
experimental data by means of a mathematical model describing a particular physical
problem. In discrete inverse theory, the mathematical model is assumed to depend on a
certain number of unknown model parameters, whose determination is the objective of the
inversion algorithm.
For the Rayleigh inverse problem, the mathematical model is given by a linear
viscoelastic, multi-layered medium. The model parameters may be the complex body wave
velocities

of

the

individual

layers,

namely

VP* = ( VP* ) 1 , ( VP* ) 2 ,... ( VP* ) n

and

VS* = ( VS* ) 1 , ( VS* ) 2 ,... ( VS* ) n . Additional sets of model parameters may include the mass

density = 1 , 2 ,... nL and the thickness h = h1 , h 2 ,... h nL

] of the layers. Because the

number of layers nL is generally assumed, the inverted profile of model parameters will
depend on the a priori assumption about nL, and it may contains large discontinuities or
other features that are not essential for matching the experimental data. By enforcing
maximum smoothness and regularity in the solution, one minimizes its dependence upon
the assumed number of layers, and at the same time rejects solutions that are unnecessarily
complicated.

Solution of the Rayleigh Inverse Problem

111

The implementation of the smoothed least-squares inversion algorithm requires an


adequate definition of smoothness of a profile of model parameters. In a multi-layered
medium, where the variation of the model parameters is discontinuous, smoothness or its
converse roughness, is defined in terms of difference rather than differential operators (Constable
et al., 1987). For complex-valued model parameters represented by the vectors VP* and VS* ,
roughness may be defined by either one of the two following expressions:
2

R 1 = V*

R2 = V

( ) ( V )

= V*
2
2

= V
2

where = P , S , the symbol

()

2
2

= V
2

(4.2)

) ( V )
H

R denotes the Euclidean norm of a vector in C N , and

indicates the Hermetian transpose of a complex-valued matrix. Finally is an n L n L


real-valued matrix representing the two-point central finite difference operator and is given by:
0 ...

1 1
0

=
... 1 1

0 1 1

(4.3)

It can be easily shown that for a continuously varying medium and for real-valued model
parameters, the two definitions of R 1 and R 2 given by Eq. 4.2 correspond to the integral
over depth of the square of the first and the second derivative, respectively, of the model
parameter function V ( y ) with respect to depth.

The experimental data are a vector of complex-valued Rayleigh phase velocities

measured at different frequencies VR* = ( VR* ) 1 , ( VR* ) 2 ,... ( VR* ) n . For the time being it will
F

be assumed the Rayleigh phase velocities refer to a specific mode of propagation (i.e., modal
Rayleigh phase velocities). However, for simplicity of notation the mode index is omitted in
the expression for VR* .
From a statistical point of view, the experimental measurements are assumed to be
independent and normally distributed. Therefore each experimental datum is completely

112

Solution of the Rayleigh Inverse Problem

described by a pair of complex numbers: the expected value ( VR* ) j ( j = 1, n F ) and its standard
deviation *j which represents an estimate of the uncertainty associated with
methodology for determining ( VR* ) j and *j is presented in Chapter 5.

(V ) .
*
R

In a linear viscoelastic, multi-layered medium characterized by the model parameters VP*


and VS* , the Rayleigh phase velocities VR* associated with a specified set of frequencies can
be predicted from the solution of the non-linear Rayleigh forward problem:

VR* = VR* ( VP* , VS* )

(4.4)

For the solution of the Rayleigh inverse problem, it is convenient to expand Eq. 4.4 in a
*
*
Taylor series about an initial guess of model parameters VP0 , VS0 obtaining:
*
VR* = VR0
+ ( J *P ) V* , V* ( VP* VP0* ) + ( J *S ) V* ,V* ( VS* VS0* ) +
P0

+ o (V V
*
P

S0

P0

) + (V

* 2
P0

*
S

* 2
S0

S0

(4.5)
2

where VR* 0 is the n F 1 vector of Rayleigh phase velocities corresponding to the solution
*
*
of Eq. 4.4 with medium parameters equal to VP0 and VS0 . The terms

(J )
*
S

VP* 0 , VS*0

(J )

*
P V * ,V *
P0
S0

and

are the n F n L complex-valued Jacobian matrices whose elements are defined by:

[( J ) ]
*
P jk

[( J ) ]
*
S jk

VP* 0 , VS*0

VP* 0 , VS*0

( VR* )
j

=
*
( VP ) k

VP* 0 , VS*0
( VR* )
j

=
*
( VS ) k

VP* 0 , VS*0

(4.6)

Solution of the Rayleigh Inverse Problem

113

where j = 1, n F and k = 1, nL . The subscripts outside the brackets indicate the point in
C n L nL at which the Jacobian matrices are evaluated. By neglecting terms higher than the
first order in Eq. 4.5, this equation reduces to:
VR* = VR0* + ( J *P ) V* , V* ( VP* VP0* ) + ( J *S ) V* , V* ( VS* VS0* )
P0

S0

P0

(4.7)

S0

*
*
*
*
which is the linearization of the functional relationship VR = VR VP , VS about the initial
*
P0

*
S0 .

model V , V

In Section 3.5.1 it was shown that the Rayleigh phase velocity is relatively insensitive to
changes in P-wave velocity in elastic media. It can be easily proved that this result also holds
in viscoelastic media for complex-valued velocities. Based on this observation and noting
that the ill-posedness of the Rayleigh inverse problem can be reduced by minimizing the
number of independent model parameters, Eq. 4.7 is inverted only for VS* . With this
assumption Eq. 4.7 simplifies to:

(J )

*
S V*
S0

VS* = ( J *S ) V* VS0* + ( VR* VR0* )


S0

(4.8)

By setting VR* = VR* , Eq. 4.8 can be used as a basis for determining the unknown profile
of model parameters VS* that correspond to a set of experimental data VR* . The value
obtained for VS* , say VS1* , could then be used as a new starting model for determining the
next approximation VS2* .
The process could then be repeated in an iterative fashion to generate a sequence of
model parameters VS0* , VS1* , VS2* . VSn* , which under suitable conditions will converge
towards the solution of the non-linear Rayleigh inverse problem. A necessary condition for
the convergence of this iterative scheme is a starting model VS0* that is sufficiently close to
the true solution.
For each iteration, the technique used to solve the linear system of equations is Occams
algorithm whose strategy is as follows: given a set of nF measured Rayleigh phase velocities

(V ) and their associated uncertainties (j = 1, n ) , find those values of


(V ) (k = 1, n ) that minimize the roughness R (or R ) of the resulting complex-valued
*
R

*
j

*
S k

114

Solution of the Rayleigh Inverse Problem

shear wave velocity profile while predicting the experimental ( VR* ) j with an acceptable
accuracy.
A measure 2 of the misfit between measured and predicted Rayleigh phase velocities
can be obtained with the weighted least-squares criterion applied to complex-valued data:

] [

2 = W* VR* W* VR* ( VS* ) W* VR* W* VR* ( VS* )


H

(4.9)

where W * is a complex-valued diagonal n L n L matrix defined by:

W* = diag 1 / 1* , 1 / *2 ,..... ,1 / *nL

(4.10)

which are the uncertainties associated with the experimental data VR* .
A standard procedure used to solve constrained optimization problems is the method of
Lagrange multipliers (Constable et al., 1987; Logan, 1997). In this case, the optimization
problem consists of minimizing the functional R 1 (or R 2 ) defined by Eq. 4.2 subject to the
condition that the residual error function 2 given by Eq. 4.9 be equal to 2* , a value
considered acceptable considering the uncertainties associated with VR* . The method of
Lagrange multipliers allows the solution of this constrained minimization problem to be
determined by finding the minimum of the following unconstrained functional:

FU = ( VS* ) ( VS* ) +
+

{[W V W V ( V )]
*

*
R

*
R

*
S

W V W V (V
*

*
R

*
R

*
S

)] }

(4.11)

2
*

where the first term is the roughness of the solution VS* , and the second term is the data
misfit multiplied by the Lagrange multiplier 1 . From Eq. 4.11 it can be observed that the
parameter may be interpreted as a smoothing parameter: If is large, the value of the
functional FU is controlled by the roughness of the solution VS* and the data misfit does
not significantly affect the solution. Conversely, if is small, most of the contribution to
FU is given by the data misfit and the roughness term plays only a minor role.

Solution of the Rayleigh Inverse Problem

115

Equation 4.11 is non-linear in VS* because of the presence of the term VR* ( VS* ) that
represents the predicted Rayleigh phase velocity. If this term is replaced by its linear part
which is given by Eq. 4.7 (without the contribution of VP* ), Eq. 4.11 becomes:

( ) ( V ) +

FU = VS*

*
S

+ W*d0* W* J*S

( )

( )

V W*d0* W* J*S
VS*0

*
S

(4.12)

V *2
*
VS 0

*
S

A necessary condition for the existence of a minimum of the unconstrained linear


functional of Eq. 4.12 is the vanishing of its gradient V * FU with the respect to VS* , that is:
S

V* FU = V* ( VS* ) ( VS* ) +
S

+ V* W* d0* W* ( J *S ) V* VS*
S
S0

] [W d W ( J )
H

*
0

*
S V*
S0

V =0

*
S

(4.13)

2
*

which can be expanded to yield:

( ) (

( ) ( )

* V* H T V* + * V*
VS S
S
VS S

+ 2 V* W*d0* W* JS*
S

( )

VS*0

V* +

( )

V W d W J

*
S

*
0

*
S V*
S0

V = 0

(4.14)

*
S

After considering that V * ( J *S ) V * = V * ( W * ) = 0 , Eq. 4.14 simplifies to:


S

T
1
*
*
+ W JS

( )

S0

( )

W* J*
S

VS*0

V* = 1 W* J*
S
*
S
VS 0

which can finally be solved for VS* to yield:

( )

W*d*
0

VS*0

(4.15)

116

Solution of the Rayleigh Inverse Problem

V = T + W* JS*

*
S

( )

( )

VS*0

( )

W* JS*

VS*0

( )

W* J*S

W*d*
0
VS*0

(4.16)

The smoothing parameter in Eq. 4.16 must be determined with the additional
constraint that the specified residual error 2* is matched with a vector VS* composed only
of negative imaginary parts. This condition will insure that the hysteretic shear-damping
ratio obtained from the inversion algorithm will be a positive quantity.
Figure 4.2 shows a flow chart of Occams algorithm applied to the solution of the
coupled Rayleigh inverse problem. Most of the computational effort of the algorithm is
spent in the solution of the complex Rayleigh eigenproblem.
Thus far, the application of Occams algorithm to the solution of the Rayleigh inverse
problem is based on comparing the experimental Rayleigh phase velocity with the predicted
Rayleigh phase velocity of a specific mode of propagation. However, the procedure will
remain valid when the modal quantities are replaced with the corresponding effective
quantities. In this case, since the effective Rayleigh phase velocity depends on two
independent variables, namely the frequency and distance from the source, the total number
of experimental data is not n F but rather n T = n F + n P , where n P is the number of
locations at which the Rayleigh phase velocity has been measured. With few other changes,
most of the formalism developed in this section for the modal Occams inversion is applicable
to the effective Occams inversion. As shown in the following sections, this generalization can
also be extended to other formulations of the Rayleigh inverse problem where different
types of analyses (coupled versus uncoupled) and response functions are considered.
In Chapter 2 it was shown that the body wave velocities ( VP* ) k , and ( VS* ) k

(k = 1, n )
L

are frequency dependent in a linear viscoelastic medium due to material dispersion. A


precise definition of the frequency dependence is given by Eqs. 2.34 and 2.35. In the
solution of the Rayleigh inverse problem presented in this section, no specific assumption
was made about the frequency-dependence law of the complex-valued vector VS* . A
technique that is often used when accounting for material is to conduct the inversion (a socalled causal inversion) at a reference frequency ref (Lee and Solomon, 1979; Herrmann,
1994). This technique is adopted in this study, and the vector VS* can be denoted by
VS*( ref ) = VS* ( ref ) . Occams algorithm remains valid; however, the partial derivatives

appearing in Eq. 4.6 must now be computed with respect to VP*( ref )
the chain rule of calculus one obtains:

and VS*( ref ) . Using


k

Solution of the Rayleigh Inverse Problem

117

INPUT
*
R

VS0*

and

Hisada -Lai
FORTRAN
Code

SOLUTION OF THE COMPLEX RAYLEIGH EIGENPROBLEM

Start

Find Roots of the


Dispersion
Equation

Compute
Eigenfunctions

Stop

Compute Greens
Function

Compute Partial
Derivatives

Compute Effective
Velocity & Partial
Derivatives

( )
( )

V R*

V S*

MATLAB 5
Program
Interface

Main Code
(Input/Output/Storage)

OCCAMS INVERSION ALGORITHM

(V

*
S ( i +1 )

MATLAB 5
Code

VS*( i ) < Tol

YES

STOP

Select a New Profile

VS*( i +1 ) = VS*( i ) + VS*( i )

NO
Solution of Linearized Inverse Problem
H
T
*
1
*
*
*
*
+ W ( J S ) V * W ( J S ) V * VS =

S0
S0

= 1 W * ( J *S ) V * W * d 0*

S0

Figure 4.2

Flow-Chart of Rayleigh Simultaneous Inversion Using Occams Algorithm

118

Solution of the Rayleigh Inverse Problem


( V * )
R j

V*
P ( ref )

( VR* )
j

=
*

( VP )
k *
VP* 0 ( ref ) , VS*0 ( ref )
k
VP 0 , VS*0

( V * )
P k

V*

P( ref )

( V * )
( VR* )
R j
j

=
*
*
V

V
(
)
S k
S ( ref )

VP*0 , VS*0
VP*0 ( ref ) , VS*0 ( ref )
k

V*
( S )k

V*

S ( ref )

(4.17)

where j = 1, n F . Finally from Eqs. 2.34 and 2.35:

( )

V*

1
k

=
2D
V*
ref
( ref )

ln
1 +

(4.18)

where = P , S .
4.6 Uncoupled Inversion
4.6.1

Overview

This section will describe several algorithms that were developed for the solution of the
uncoupled Rayleigh inverse problem. Uncoupled inversion refers to the case where the
dispersion and the attenuation curves are inverted independently. For weakly dissipative
media the uncoupled analysis yields satisfactory results although, as mentioned in Section
4.3, the uncoupled inversion is more ill-posed than the corresponding coupled analysis.
Figure 4.3 shows classes of possible algorithms for solving the uncoupled Rayleigh
inverse problem. The shaded boxes indicate algorithms developed during this study,
whereas the boxes with dashed borders briefly outline the procedure used to obtain the
experimental data. Each class of algorithm has two components: one to compute the shear
wave velocity profile ( VS ) k ( k = 1, nL ) and the other to determine the shear damping ratio

profile (DS ) k . The subscript eff appearing in the algorithms UEQMA and UEFMA is used
to denote effective quantities. The term G in the expression R = G D S denotes the matrix
formed by the partial derivatives of Rayleigh phase velocity with respect to the shear and
compression wave velocities of the soil layers (hereafter called the G-matrix). This matrix is
calculated using Eq. 3.63b where the partial derivatives are computed using Eq. 3.45.

Uncoupled Inversion
Fundamental Mode Analysis
[UFUMA]
Shear Wave Velocity
Inversion
Non-Linear Inversion
VR = VR(Vs)
[Compute VR using first mode only]

Experimentally
obtain VR() from
linear regression
arg[w(r,)] = a + k R r

Equivalent Multi-Mode Analysis


[UEQMA]

Shear Damping Ratio


Inversion
Linear Inversion
R = GDs
[Compute G using first mode only]

Experimentally
obtain R() from
non-linear regression
|w(r,)| = Fy Gy(r,)exp(- R r)
Non-Linear Inversion
|w(r,)| = Fy Gy (r,) exp(-GDs r)
[Account for R = R(r) = GDs]

Shear Wave Velocity


Inversion

Shear Damping Ratio


Inversion

Non-Linear Inversion
VR(eff) = VR(eff)(Vs)

Linear Inversion
R(eff) = G(eff) Ds

[Ignore dependenceVR(eff) =VR(eff)(r)]

Experimentally
obtain VR(eff)() from
linear regression
arg[w(r,)] = a + kR(eff) r

[Ignore dependence G(eff) = G(eff)(r)]

Experimentally
obtain R(eff)() from
non-linear regression
|w(r,)| = Fy Gy(r,)exp(-R(eff) r)

Effective Multi-Mode Analysis


[UEFMA]
Shear Wave Velocity
Inversion
Non-Linear Inversion
VR(eff) = VR(eff)(Vs)
[Account for VR(eff) = VR(eff)(r)]

Experimentally
obtain VR(eff)(, r) from
VR(eff)(, r) = r/
at each receiver spacing
Displacement Phase Spectra
Non-Linear Inversion
arg[w(r,)] = (Vs)

Experimentally
obtain arg[w(r,)]
at each receiver location
and at various

Figure 4.3 Algorithms for the Solution of the Uncoupled Rayleigh Inverse Problem

Shear Damping Ratio


Inversion
Non-Linear Inversion
|w(r,)| = Fy Gy(r,) exp(-G(eff)Ds r)
[Accounts for G(eff) = G(eff)(r)]

Experimentally
obtain |w(r,)|
at each receiver location
and at various

120

Solution of the Rayleigh Inverse Problem

In Eq. 3.63b the compression-damping ratio D P has been replaced by the term K D S
where K is a parameter defining the ratio of compression to shear damping ratio (Rix et al.,
1998a). Previous studies (Spang, 1995) have shown that the value adopted for K has a
negligible influence on the backcalculated shear damping ratio profile. This is due to the fact
that the phase velocity of Rayleigh waves is relatively insensitive to changes in VP and thus
the partial derivative VR VP in Eq. 3.63b is small compared to VR VS (see Section
3.5.1). Furthermore, many studies (Winkler and Nur, 1979; Jongmans, 1990; Malagnini,
1996; Leurer, 1997) have shown that 0 < K < 1 in soil. This fact (which is consistent with
the observation that, in fine grained materials such as soils, most of the energy dissipation is
expected to occur in shear mode) combined with small values of VR VP makes the
second term of Eq. 3.63b negligible compared to the first. Based on these considerations, all
uncoupled analyses were performed using a value of K = 1.
In an uncoupled analysis of Rayleigh wave data, the inversion of the attenuation curve,
that is R = G D S , is a linear problem. In this case it can be shown that the formulation of
Occams algorithm could be reversed and would yield the same result. In other words, for a
linear inversion the smoothest solution with specified error misfit is the same as the solution
where the error misfit is minimized for a specified value of smoothness. The Lagrange
multipliers of the two solutions are reciprocals of each other.
The following sections provide a description of the main tasks performed by each class
of algorithms. The inversion algorithms were written in MATLAB; the Rayleigh forward
problem was written in FORTRAN 77 and linked to the main program via the MATLAB
API (Application Program Interface). A description of the main tasks performed by the
algorithms is reported in Appendix C.
4.6.2

Uncoupled Fundamental Mode Analysis

This is the simplest type of uncoupled inversion analysis and is based on independently
comparing the experimental dispersion and attenuation curves with the corresponding
curves obtained from the theoretical model for the fundamental mode of Rayleigh wave
propagation. The algorithm is composed of two modules called Dispersion and Attenuation.
The Dispersion module performs a non-linear inversion of the experimental dispersion curve
using Occams algorithm (Eq. 4.16). In performing this task Dispersion calls a MATLAB
mex-file called Rayleigh, which is a FORTRAN 77 routine that solves the eigenvalue problem
of Rayleigh waves in elastic, vertically heterogeneous media (modified from Hisada, 1995).
Rayleigh also computes the partial derivatives of the modal Rayleigh phase velocities with
respect to the shear and compression wave velocities of the soil layers using the variational
formulation. Rayleigh computes the theoretical dispersion curve and the partial derivatives of
Rayleigh phase velocity with respect to medium parameters for the fundamental mode of
Rayleigh wave propagation. The experimental phase velocities are imported into Dispersion
from an external ASCII file.

Solution of the Rayleigh Inverse Problem

121

The second module of the UFUMA algorithm is called Attenuation and is designed to
implement the linear inversion of the experimental attenuation curve. This task is
accomplished with a two-step procedure. First, the frequency-dependent Rayleigh
attenuation coefficients R ( ) are computed from the experimental displacement
amplitudes at multiple receiver offsets using a non-linear regression based on the expression
w(r , ) exp. = Fy Gy (r , ) e R r where w(r, ) is the vertical particle displacement spectrum,
and Gy (r , ) is the vertical geometric spreading function. Once the experimental
attenuation curve is determined, the shear damping ratio profile is determined from the
linear inversion of the relation R = G D S . Both Gy (r , ) and G and calculated by
Rayleigh.
4.6.3

Uncoupled Equivalent Multi-Mode Analysis

This class of inversion algorithms explicitly recognizes the multi-mode nature of the
quantities measured in surface wave tests. Accordingly, during the inversion process the
experimental dispersion and attenuation curves are matched with the effective theoretical
dispersion and attenuation curves instead of the modal curves. However, since the effective
Rayleigh phase velocities V$ R ( , r ) and attenuation coefficients $ R ( , r ) are functions of
two independent variables (frequency and source-receiver distance), a simplification is
required to reduce the dimensionality of the problem by one, that is to transform dispersion
and attenuation surfaces into equivalent curves. The simplification is suggested by the
experimental technique used for the uncoupled phase velocity and attenuation
measurements and involves an averaging process of V$ R ( , r ) and $ R ( , r ) over the
receiver spacings used in the measurements.
The UEQMA algorithms UEQMA are composed of two modules that are again called
Dispersion and Attenuation. The Dispersion module implements the non-linear inversion of the
experimental dispersion curve using Occams algorithm. Like UFUMA, Dispersion calls a
MATLAB mex-file named Rayleigh that solves the eigenvalue problem of Rayleigh waves in
elastic, vertically heterogeneous media. Rayleigh also computes the effective Rayleigh phase
velocity V$ R ( , r ) , and the partial derivatives of V$ R ( , r ) with respect to the shear and
compression wave velocities of the soil layers. Dispersion computes the effective theoretical
dispersion curve and the effective partial derivatives that account for all of the Rayleigh
modes of propagation according to Eqs. 3.23 and 3.58, respectively. Averaging the effective
Rayleigh phase velocities and partial derivatives over the receiver offsets used in the
measurements eliminates the dependence of these quantities on the spatial coordinate. The
experimental dispersion curve and receiver offsets are imported into Dispersion from an
external ASCII file.
Attenuation is the second module of the algorithm UEQMA and is used to obtain the
shear damping ratio profile from the experimental attenuation curve. Like UFUMA, the

122

Solution of the Rayleigh Inverse Problem

inversion of the experimental attenuation curve is performed using a two-step procedure


that involves computation of the attenuation coefficients R ( ) first, followed by an
inversion of the R ( ) yielding the D S profile. The key difference is that the linear
inversion R = G D S is performed with the effective rather than the modal G-matrix.
4.6.4

Uncoupled Effective Multi-Mode Analysis

UEFMA denotes a class of inversion algorithms that take the multi-mode nature of
Rayleigh wave propagation in vertically heterogeneous media, fully into account in the
interpretation of surface wave tests. Ideally, the experimental measurements would consist
of a series of phase velocities V$ R ( , r ) and a series of attenuation coefficients $ R ( , r )
each measured at different frequencies and at different receiver locations. These two
independent sets of data constitute what may be called the experimental dispersion and
attenuation surfaces of a site. The UEFMA algorithm then finds the smoothest shear wave
velocity and shear-damping ratio profiles whose corresponding theoretical dispersion and
attenuation surfaces match those obtained from the experiments.
Unfortunately, the practical implementation of this algorithm is difficult because of the
problems associated with the experimental determination of V$ R ( , r ) and $ R ( , r ) at a
point. The local (i.e., spatially dependent) nature of these quantities suggests that they be
determined from displacement phase and amplitude measurements performed over very
small receiver spacings. However, the uncertainty associated with these measurements will
increase as the distance between the receivers decreases. This problem becomes particularly
acute for displacement amplitude measurements because of the combined effect of
geometric and material attenuation.
In Chapter 5 it is shown that the dispersion surface can be constructed by
differentiating the displacement phase with respect to the source-to-receiver distance at
constant frequency. The attenuation surface is similarly determined from the displacement
amplitude. In both cases, the procedure used to determine the Rayleigh phase velocity and
the attenuation coefficient at different receiver offsets is numerical differentiation, which is an
ill-conditioned problem particularly with inaccurate or noisy data.
A possible strategy to overcome the difficulties of determining the experimental
dispersion and attenuation surfaces is to apply a numerical fitting procedure to the set of
data formed by the experimental phases and displacement amplitudes. This approach will
enable obtaining smooth phase and corrected amplitude surfaces over which the operation of
differentiation could be performed analytically. This requires the selection of an adequate
model for the numerical fitting procedure, which may be guided by the analytical
expressions obtained in Chapter 3 for the displacement phase and amplitude.
The difficulties that are inherent in the determination of the experimental dispersion
and attenuation surfaces suggest a new approach to interpreting surface wave data. As

Solution of the Rayleigh Inverse Problem

123

mentioned at the beginning of this chapter, the conventional interpretation of surface wave
measurements is based on adopting dispersion and attenuation curves as the medium
response functions. This choice is not unique, and other types of response functions may be
selected that may be better suited to the solution of the Rayleigh inverse problem. In the
frequency domain some of these alternate response functions include displacement
amplitude, displacement phase, and displacement spectra (see Fig. 4.1).
In surface wave tests displacement phase and amplitude are directly measured; Rayleigh
phase velocity and attenuation coefficient are derived from the phase and amplitude. Thus,
it may be advantageous to choose the displacement phase and amplitude as the response
functions for the solution of the Rayleigh inverse problem because they are more basic
quantities. For the uncoupled inversion an example of this approach would be to determine
the shear damping ratio profile directly from the experimental displacement amplitudes
$
using w(r , ) exp. = Fy Gy (r , ) eGDS r . In essence, the response function is changed from
attenuation coefficients to displacement amplitudes. This alternative, denoted in Fig.4.3 with
a dotted box, is attractive because it permits the experimental displacement amplitudes to be
used directly and avoids the need to calculate the attenuation coefficients.
A direct inversion of the measured displacement phases (,r ) would be the
analogous procedure for determining the shear wave velocity profile. However, the
inversion of the displacement phases is far more involved than the inversion of the
displacement amplitudes, mainly because the functional relationship between the
displacement phase with the shear wave velocity profile does not admit a simple, explicit
representation. Moreover, the partial derivatives of the displacement phase with respect to
the medium parameters, required for the solution of this non-linear inverse problem, can
only be computed numerically. Section 4.7.4 presents a unifying approach for the solution
of the Rayleigh inverse problem where the displacement phases and amplitudes are replaced
by complex displacement spectra.
4.7 Coupled Inversion
4.7.1

Overview

In this study, the term strongly coupled inversion describes a procedure in which the
dispersion and attenuation curves are inverted simultaneously. A strongly coupled Rayleigh
inversion requires the ability to solve the complex eigenproblem for linear viscoelastic media
where Rayleigh phase velocity VR ( ) and attenuation coefficient R ( ) both depend on

the body wave velocities, VP ( y) and VS ( y ) , and the material damping ratios, D P ( y ) and
D S ( y) . In other words, a strongly coupled inversion refers to any procedure where an

124

Solution of the Rayleigh Inverse Problem

appropriate complex-valued response function R * = R

(V

*
P

, VS* ) is inverted to obtain the

corresponding medium parameters VP* , and VS* .


A consistent, strongly coupled inversion should not only simultaneously invert the real
and the imaginary parts (or amplitude and phase) of R * = R

(V

*
P

, VS* ) , but it should also

include an experimental procedure where both (R * ) and (R * ) are measured


simultaneously. In other words, in a consistent, coupled inversion the dispersion and
attenuation curves should be both measured and inverted simultaneously. Chapter 5 will
illustrate the details of an experimental procedure conceived for the simultaneous
measurement of both VR ( ) and R ( ) .
In the seismological literature the term coupled inversion is often used for a procedure
that accounts for the link between surface wave phase velocity and attenuation and the
stiffness and damping properties of the layers in weakly dissipative medium (Aki and
Richards, 1980; Keilis-Borok, 1989; Herrmann, 1994). As illustrated in Section 3.5.3, this
weakly coupled inversion requires only the solution of the real-valued eigenproblem in elastic
media. The implementation of a weakly coupled inversion is based on the iterated
application of Eq. 3.63.
Figure 4.4 shows some of the possible algorithms that may be used for the solution of
the coupled Rayleigh inverse problem. The shaded boxes show the strongly coupled
inversion algorithms developed in this study and written in MATLAB; the solution of the
corresponding forward problem for linear viscoelastic media was written in FORTRAN 77.
A description of the main tasks performed by these computer codes is reported in
Appendix C.
The dashed boxes in Fig. 4.4 briefly outline the principles used to obtain the
experimental data. The symbol T (r, ) appearing in the dashed boxes denotes the
displacement transfer function; a quantity that will be defined in Chapter 5. The term
R (r , ) = *v (r ,0 , ) denotes the complex-valued phase angle defined by Eq. 3.76.
Finally, the subscript eff appearing in some of the equations is used to distinguish effective
quantities from modal quantities. The next three sections will provide a description of the
tasks performed by each of the algorithms outlined in Fig. 4.4.

Coupled Inversion
Fundamental Mode Analysis
[CFUMA]

Equivalent Multi-Mode Analysis


[CEQMA]

Effective Multi-Mode Analysis


[CEFMA]

Shear Wave Velocity and


Shear Damping Ratio
Inversion

Shear Wave Velocity and


Shear Damping Ratio
Inversion

Shear Wave Velocity and


Shear Damping Ratio
Inversion

Non-Linear Inversion
VR = VR(Vs)

Non-Linear Inversion
VR(eff) = VR(eff)(Vs)

Non-Linear Inversion
VR(eff) = VR(eff)(Vs)

[Compute V R using first mode only]


(Note: V R and V S are complex-valued)

[Ignore dependence VR(eff) = VR(eff)(r)]


(Note: V R and V s are complex-valued)

[Account for V R(eff) = VR(eff)(r)]


(Note: V R and Vs are complex-valued)

Experimentally

Experimentally

Experimentally

obtain VR() and R()


from non-linear regression
T(r,) = G(r,) exp(-ikR r)

obtain VR(eff)() and R(eff)()


from non-linear regression
T(r,) = G(r,) exp(-ik R(eff) r)

obtain VR(eff)(r,) and R(eff)(r,)


iteratively from solving
T(r,) = G(r,) exp[-iR(r,)]

Complex Displacement Spectra


Non-Linear Inversion
w(r,) = w(Vs)
(Note: w(r,) and Vs are complex valued)

Experimentally
obtain w(r,) from
arg[w(r,)] and |w(r,)|
at each receiver location

Figure 4.4 Algorithms for the Solution of the Strongly Coupled Rayleigh Inverse Problem

126
4.7.2

Solution of the Rayleigh Inverse Problem


Coupled Fundamental Mode Analysis

This algorithm implements the simultaneous, strongly coupled inversion of the


experimental dispersion and attenuation curves by comparing these curves with the
theoretical curves obtained from the viscoelastic model for the fundamental mode of
propagation. The algorithm CFUMA performs in a single step a task that is performed by
the algorithm UFUMA in two separate operations with the additional, beneficial internal
constraint provided by the Cauchy-Riemann equations. The program that implements
CFUMA is a MATLAB code called ViscoRay, which performs the simultaneous inversion of
the experimental dispersion and attenuation curves by applying the complex formalism to a
constrained least squares algorithm that enforces maximum smoothness in the inverted
complex shear wave velocity profile.
ViscoRay uses a MATLAB mex-file called Rayleigh written in FORTRAN 77 that solves
the complex eigenvalue problem of Rayleigh waves in linear viscoelastic multi-layered media.
Rayleigh also computes the partial derivatives of the complex Rayleigh phase velocity with
respect to the complex shear and compression wave velocities of the soil layers using the
variational formulation illustrated in Section 3.8. These partial derivatives are used to
construct the complex-valued Jacobian matrix required for the solution of the coupled
Rayleigh inverse problem.
The theoretical dispersion and attenuation curves computed by Rayleigh include only the
fundamental mode of propagation of Rayleigh waves. However, the frequency dependent
attenuation coefficients R ( ) are calculated using a geometric spreading function
Gyv (r , ) that accounts for all of the modes of propagation (Rix et al., 1998a). The
experimental Rayleigh phase velocities and attenuation coefficients are imported into
ViscoRay from external ASCII files.
4.7.3

Coupled Equivalent Multi-Mode Analysis

This algorithm simultaneously inverts the experimental dispersion and attenuation


curves by matching them with the corresponding effective, theoretical curves obtained from
the viscoelastic model. As in UEQMA, the effective theoretical dispersion and attenuation
curves are obtained from the respective surfaces by averaging the complex effective phase
velocity V$ R* ( , r ) over the receiver spacings used in the actual test. CEQMA is
implemented in MATLAB code named ViscoRay that performs the simultaneous inversion
of the experimental complex dispersion curve using Occams algorithm.
ViscoRay uses a MATLAB mex-file called Rayleigh written in FORTRAN 77 that solves
the complex eigenvalue problem of Rayleigh waves in viscoelastic multi-layered media.
Rayleigh computes the effective complex Rayleigh phase velocity and its partial derivatives
with respect to the complex shear and compression wave velocities of the medium layers
using the variational formulation of Section 3.8.

Solution of the Rayleigh Inverse Problem

127

The effective theoretical dispersion and attenuation curves and the effective partial
derivatives of Rayleigh phase velocity with respect to medium parameters computed by
Rayleigh, reflect the contribution of all the Rayleigh modes of propagation. It should be
noted that the experimental attenuation coefficients are calculated iteratively using the
geometric spreading function Gyv (r , ) that depends on the shear wave velocity profile of
the current iteration.
4.7.4

Coupled Effective Multi-Mode Analysis

This is the most sophisticated algorithm presented in this study for the inversion of
surface wave data. It combines the features of the simultaneous inversion with a thorough
consideration of the multi-mode nature of Rayleigh wave propagation in multi-layered
media, which is reflected in the spatial dependence of the effective phase velocity and
attenuation coefficient. Unfortunately, most of the difficulties associated with the
experimental determination of the dispersion and attenuation surfaces discussed in Section
4.6.4 in the context of UEFMA also apply to CEFMA.
As it will be shown in Chapter 5, the difficulties of the algorithm CEFMA are related to
the experimental determination of the effective wavenumber, which is obtain from an
unstable process of numerical differentiation of the complex phase angle *v (r ,0 , ) with
respect to the source-to-receiver distance. As mentioned in section 4.6.4, a smooth
numerical fitting of the experimental data, in this case the displacement transfer functions
T (r, ) , may mitigate some of these difficulties. Once the experimental dispersion surface
V$ * ( , r )
has been defined, the objective of the algorithm CEFMA involves the task of

exp.

finding the smoothest complex shear wave velocity profile VS* ( ) whose theoretical
complex dispersion surface matches V$ R* ( , r ) exp. .

As shown in Fig.4.4, the algorithm CEFMA can be implemented following a different


strategy. Because the quantity measured experimentally is the displacement transfer function
T (r, ) , it follows that the most natural inversion of surface waves data should involve this
response function. The effective phase velocity V$ R* ( , r ) is obtained from T (r, ) via a

roughing filter operation, and hence the inversion of V$ R* ( , r ) data is necessarily an illconditioned operation. The inversion of surface waves data based on the use of
displacement response functions is not a common practice to this day.
However, in the opinion of the writer the direct inversion of displacement spectra
w(r, ) or transfer functions T (r, ) constitutes the most rational approach to the
interpretation of active surface waves tests. The major difficulty of its practical
implementation is the computation of the partial derivatives of the displacement spectra

128

Solution of the Rayleigh Inverse Problem

with respect to the medium parameters. Currently, these partial derivatives have to be
computed numerically because no closed-form solutions, functions of the unperturbed
medium parameters, are available. Recent works in theoretical seismology are attempting to
address the solution of this difficult problem (Zeng and Anderson, 1995).

5 RAYLEIGH PHASE VELOCITY AND ATTENUATION


MEASUREMENTS

5.1 Overview
The most important aspects of surface wave measurements are reviewed in this chapter.
Surface wave tests are often called Spectral-Analysis-of-Surface-Waves (SASW) tests in the
engineering literature (Nazarian, 1984; Stokoe et al., 1989; Tokimatsu, 1995). Although
surface waves include both Rayleigh and Love waves, most of the methods currently used to
near-surface soil and rock properties focus exclusively on the observation of Rayleigh waves.
The SASW test is a relatively young in-situ seismic technique (compared to more
traditional seismic tests such as cross-hole and down-hole tests) that was developed by
Stokoe and co-workers during the early 1980s (Nazarian et al., 1983; Nazarian, 1984;
Snchez-Salinero, 1987; Rix, 1988; Stokoe et al., 1989). The technique evolved from the
Steady-State-Vibration Technique used by the US Army Corps of Engineers Waterways
Experiment Station (WES) during the early 1960s (Richart et al., 1970). The SASW test is
becoming increasing popular in the geotechnical engineering community, primarily because
it is a non-invasive field technique, and hence does not require the use of boreholes or
probes. This attractive feature may be crucial for certain types of geotechnical investigations.
Surface wave tests were originally developed for the determination of the elastic moduli
profile of soil deposits and pavement systems (Nazarian et al., 1983; Stokoe et al., 1989).
More recently, Rix et al. (1998a) developed a technique for using the SASW test to
determine the material damping ratio profile of a layered soil deposit. So far, however, the
two problems of determining the stiffness and material damping ratio profiles of a site have
been treated separately (i.e., uncoupled). Section 4.7 described several algorithms for the
simultaneous (i.e., strongly coupled) inversion of surface wave dispersion and attenuation
data. Section 4.7 also noted that the dispersion and attenuation data should not only be
inverted but also measured simultaneously. One goal of this chapter is to illustrate an
experimental procedure for the simultaneous measurement of both dispersion and
attenuation data.
Chapter 5 is organized in two main sections: Section 5.2 reviews the methods that are
currently used in conventional surface wave measurements. Section 5.3 proposes a new
approach to surface wave measurements where consistency between measurement
techniques and data interpretation is emphasized. Echoing the inversion analyses of Chapter
4, the new methodology is developed following two strategies. For the uncoupled inversion
algorithms presented in Section 4.6, the dispersion and attenuation data should be obtained
from uncoupled measurements, a topic that is discussed in Section 5.3.1. Conversely, the
use of the simultaneous inversion algorithms discussed in Section 4.7 suggests that the
129

130

Rayleigh Phase Velocity and Attenuation Measurements

dispersion and attenuation data be measured simultaneously. Coupled measurements of


surface wave data are discussed in Section 5.3.2. The chapter concludes with some
considerations concerning the statistical errors associated with surface wave measurements.
5.2 Conventional Measurements Techniques
Surface wave methods are traditionally divided into active and passive methods
(Tokimatsu, 1995). In the active methods Rayleigh waves are generated by either an
impulsive or a vertically oscillating harmonic source applied at the free surface of a vertically
heterogeneous medium. The ensuing particle motion is recorded by an array of receivers
placed on the ground surface in line with the source. Since active surface wave methods
have a penetration depth that is typically on the order of 15 to 20 meters, they are well
suited for near-surface site characterization. The major obstacle to greater penetration
depths is the difficulty of generating lower frequency (i.e., longer wavelength) with portable
sources.
Passive methods overcome this limitation because they do not involve generation of
wave energy with artificial sources. They are based on the observation of short- and longperiod ground motion induced by cultural noise and microtremors. Passive methods require
the particle motion to be recorded by a large number of sensors arranged in twodimensional arrays over the ground surface. Penetration depths with passive methods can
range from less than 50 meters with short-period microtremors to several kilometers with
long-period microtremors (Tokimatsu, 1995). This study focuses exclusively on the
interpretation of active surface wave methods, particularly those involving harmonic
sources.
The receivers used in surface wave tests for near-surface site characterization are usually
vertical velocity transducers with natural frequencies ranging from 1 to 4.5 Hz (Stokoe et al.,
1989). The recording device is usually a Fast Fourier Transform (FFT) dynamic signal
analyzer that is capable of performing real-time spectral analyses of the particle velocity time
histories measured at the receivers. Figure 5.1 shows a typical configuration of the
equipment used in SASW testing during Rayleigh phase velocity measurements.
Sources are classified as transient and harmonic according to the time-variation of the
dynamic force. Transient sources are typically impulsive; common examples include
sledgehammers and dropped weights. They generate surface waves containing a broad range
of frequencies. Lack of repeatability and poor signal-to-noise ratios are their most important
limitations. Common harmonic sources include hydraulic vibrators and electro-mechanical
shakers that sweep through a pre-selected range of frequencies. For SASW tests performed
in soil deposits, the frequency range is approximately 5 to 200 Hz (Rix, 1988). Harmonic
sources are more repeatable and capable of higher signal-to-noise ratios than transient
sources (Rix, 1988; Spang, 1995). Moreover, the use of harmonic sources greatly simplifies

Rayleigh Phase Velocity and Attenuation Measurements

131

Recording
Device

Source

Receivers

Vertical
Particle
Motion

Figure 5.1

Typical Configuration of the Equipment Used in SASW Testing

the interpretation of surface wave data since the analysis of Rayleigh wave propagation in
vertically heterogeneous media is more difficult for transient than for harmonic waves.
5.2.1

Phase Velocity Measurements

Figure 5.2 shows a typical source-receiver configuration used during conventional phase
velocity measurements (i.e., the two-station method). Rayleigh waves are generated by a
harmonic source oscillating at a circular frequency . Two receivers located at distances r1
and r2 from the source detect the vertical particle motion that is recorded by the signal
analyzer in the form of particle velocity spectra V (r, ) defined to be the Fourier transform
of the particle velocity w
& ( r , t ) = w t .

From the velocity spectra V (r, ) , two other important spectral quantities are
calculated: the auto-power spectrum G rr ( ) of each receiver and the cross-power spectrum
G r1r2 ( ) of the two receivers:

Grr () = V(r , ) V(r , )


Gr1r2 ( ) = V(r1 , ) V(r2 , )

(5.1)

132

Rayleigh Phase Velocity and Attenuation Measurements

Fy e it
w 1 (r1 , t )

Rayleigh Wave

w 2 ( r2 , t )

Geophones

r1
r2

Figure 5.2

Source-Receivers Configuration in SASW Phase Velocity Measurements

The symbol () in Eq. 5.1 denotes complex conjugation. The time delay between

receivers as a function of the circular frequency is given by arg G r1r2 ( ) . Hence, the
phase velocity VR ( ) of the propagating Rayleigh wave can be computed from:

VR ( ) =

(r2 r1 )

arg G r1r2 ( )

(5.2)

Equation 5.2 yields the experimental dispersion curve associated with the pair of
receivers located at r1 and r2. This procedure is then repeated for different receiver spacings,
and the individual dispersion curves are combined together to form the composite
dispersion curve of the site (Nazarian, 1984).
Receiver positions are chosen according to one of two schemes: the common source
array and the common receiver midpoint array. Both schemes are illustrated in Figs.5.3 (a)
and (b). Local stratigraphy and portability of the source control the selection of the receiver
array. Sometimes the SASW test is performed using both forward and reverse arrays
(Stokoe et al., 1989). The receiver positions remain the same, but the source is moved to the
opposite end of the array. The results of the forward and reverse arrays are combined
together in an attempt to mitigate the effects of lateral inhomogeneities and/or local
discontinuities.
In Fig.5.2 the receivers are typically spaced with a ratio r2 r1 = 2 (Snchez-Salinero,
1987). The receiver spacing should also chosen to eliminate spatial aliasing. According to the

Rayleigh Phase Velocity and Attenuation Measurements

133

Nyquists criterion, spatial aliasing is avoided if the receiver spacing is chosen to be smaller
than half the wavelength one wants to measure.
An important spectral quantity that is also calculated during SASW phase velocity
measurements is the ordinary coherence function, which is defined by:

F e i t

Receivers
2

1
F e i t

Receivers

F e i t

Receivers

Figure 5.3(a)

SASW Arrangement Using Common Receiver Midpoint Array

near field
F e i t

Receivers
2

1
F e i t

Receivers

F e i t

Figure 5.3(b)

Receivers

SASW Arrangement Using Common Source Array

134

Rayleigh Phase Velocity and Attenuation Measurements

r21r2 ( ) =

[G
[G

r1r2
r1r1

][
( )] [G

]
( ) ]

( ) Gr r ( )
12

r2 r2

(5.3)

The ordinary coherence function gives a measure of how the measured particle velocity

w
& (r1 , t) is related to w
& (r2 , t ) . It can be shown (Bendat and Piersol, 1986) that 0 2r1r2 1
with the upper bound r1r2 = 1 corresponding to a situation where there is an exact linear

relationship between w
& (r1 , t) and w
& (r2 , t ) . Low coherence values may be attributed to the
presence of signal noise or more generally to the situation where the measured particle
velocities at the receiver r1 and r2 are not linearly related. From Eq. 3.28 with
u y (r ,0 , ) = w(r , ) it can be easily shown that the theoretical coherence function (i.e., the
coherence function computed with synthetic particle velocities) is equal to one at all
frequencies. Therefore, since Eq. 3.28 has been derived by neglecting the body wave field,
an additional cause for observed low values of the coherence function may be the near-field
effects.
In the context of Section 3.3, the Rayleigh phase velocity computed with Eq. 5.2 is the
average effective Rayleigh phase velocity V$ R (r , ) over the distance (r2 r1 ) . This quantity,
which is sometimes called the apparent phase velocity (Tokimatsu, 1995), reflects the
contribution of several modes of propagation of Rayleigh waves, and its magnitude depends
on the location where it is measured as demonstrated by Eq. 3.23. Therefore, the current
procedure based on Eq. 5.2 eliminates important information from the experimental
measurements, that is the functional dependence of the measured Rayleigh phase velocity
on the receiver position. As discussed in Chapter 4, the well posedness of the Rayleigh
inverse problem can be improved by supplying more information to constrain the solution.
Thus, a substantial improvement is expected by accounting for the dependence of the
measured phase velocity on the source-to-receiver distance.
5.2.2

Attenuation Measurements

The equipment configuration used in surface wave attenuation measurements is the


same as that shown in Fig. 5.1. The common source array shown in Fig. 5.3b is typically
used. Rayleigh attenuation coefficients R ( ) are obtained from measurements of the
vertical displacement amplitudes w(r , ) at several receiver offsets over a specified range of
frequencies.
The relevant spectral quantity is now the particle velocity auto-power spectrum Grr ( )
calculated at each receiver spacing. Because ambient noise may be important, particularly at

Rayleigh Phase Velocity and Attenuation Measurements

135

large receiver offsets, the experimental particle velocity spectra are corrected to account for
the noise effects (Rix et al., 1998a; Rix et al., 1998b):

~
Grr ( ) = ~
sr2 ( ) Grr ( )

(5.4)

where Grr ( ) is the measured auto-power spectrum that is presumed to contain non-

sr2 ( ) is the ordinary coherence function between the


coherent noise. The quantity ~

sr ( )
harmonic source and the measured vertical particle velocity. Computation of ~
requires the use of an accelerometer at the source to monitor the motion of the harmonic
oscillator.
2

From Eq. 5.4 the experimental vertical particle displacement spectrum is readily
computed by:

w( r , ) =

V (r , )
C( )

G rr ( )

(5.5)

C( )

where C( ) is a frequency dependent calibration factor that converts the output of the
velocity transducer (volts) into engineering units (e.g., cm/sec).
Once the vertical displacement amplitudes w (r, ) have been computed, the Rayleigh

attenuation coefficients R ( ) can be determined from Eq. 3.76 by assuming


yv* (r ,0 , ) K * ( ) r

where

K * ( ) = VR ( ) + i R ( ) .

This

assumption

is

equivalent to assuming that the complex phase angle yv* (r ,0 , ) is dominated by the
fundamental mode of propagation. This hypothesis applies only to the phase angle and not
to the modulus of the vertical displacement w(r, ) . In other words, the attenuation
coefficients are calculated using a hybrid approach in which the material attenuation is
assumed to be dominated by the fundamental mode, but the geometric attenuation is
correctly computed by accounting for all the modes of propagation. With these assumptions
Eq. 3.76 gives:
w(r , ) = Fy G (r , ) e R ( )r

(5.6)

136

Rayleigh Phase Velocity and Attenuation Measurements

where w(r , ) = u y (r ,0 , ) and G (r , ) = Gyv (r ,0 , ) . Equation 5.6 forms the basis of a


non-linear regression analysis to determine the frequency-dependent attenuation coefficients
R ( ) from the experimental displacement amplitudes w(r, ) . The geometric spreading

function G (r, ) can be computed from Eq. 3.25b since G (r , ) = U$y* (r ,0 , ) . For weakly
dissipative media U$y* (r ,0 , ) U$y (r ,0 , ) and G (r, ) can be computed from the elastic
solution of the Rayleigh forward problem.
Figures 5.4(a) and (b) show the results of two experimental attenuation coefficients
R ( ) computed at the Treasure Island National Geotechnical Experimentation Site
(NGES) at two different frequencies. The magnitude of the harmonic force Fy may be
considered as an additional parameter to be determined from the regression. In this case the
two parameters R ( ) and Fy may be determined using a partitioned non-linear regression
algorithm that eliminates the linear parameter Fy from the non-linear regression for

Vertical Particle Displacement (m)

R ( ) . It can be shown (Lawton and Sylvestre, 1971) that this procedure is beneficial in
accelerating and stabilizing the convergence of the algorithm.
1.E-05
Experimental
Theoretical

f = 30 Hz
R = 0.0377 1/m

1.E-06

1.E-07

1.E-08

1.E-09
0

10

20

30

40

50

Distance (m)
Figure 5.4(a)

Attenuation Coefficient Computation at Treasure Island Site

60

Vertical Particle Displacement (m)

Rayleigh Phase Velocity and Attenuation Measurements

137

1.E-05
Experimental
Theoretical

f = 68 Hz
R = 0.0236 1/m

1.E-06

1.E-07

1.E-08

1.E-09
0

10

20

30

40

50

60

Distance (m)
Figure 5.4(b)

Attenuation Coefficient Computation at Treasure Island Site

Alternatively, the dynamic force Fy can be calculated by measuring the acceleration of


the harmonic oscillator, and hence it can be introduced in Eq. 5.6 as a known term; it
should be noted that in general Fy = Fy ( ) .
Obviously, the procedure of computing the R ( ) by specifying the force magnitude
Fy = Fy ( ) is preferred. The experimental attenuation coefficients R ( ) determined by

Eq. 5.6 are average values of the effective attenuation coefficient $ R (r , ) over the receiver

locations. As in the case of phase velocity measurements, the dependence of $ R (r , ) on


the source-to-receiver distance is eliminated via this averaging procedure. However, the
averaging procedure differs for V$ R (r , ) and $ R (r , ) .
In phase velocity measurements, a two-station method is used to determine VR ( ) for
various receiver spacings. The experimental composite dispersion curve is then obtained by
averaging the individual dispersion curves obtained at each receiver spacing. A two-station
method applied to attenuation measurements, although in principle possible, would likely
yield inaccurate results because of the combined effect of material and geometric
attenuation. Over short receiver-to-receiver distances such as those commonly used in
surface wave testing, geometric attenuation controls the spatial decay of surface waves
generated by point sources; material attenuation is difficult to extract. A multi-station

138

Rayleigh Phase Velocity and Attenuation Measurements

method overcomes this problem because material attenuation is more easily observed and
measured experimentally over larger distances.
The averaging procedure involved in the determination of the attenuation coefficient
R ( ) is based on the assumption that yv* (r ,0 , ) K * ( ) r in Eq. 3.76. This

corresponds to using a lumped attenuation coefficient R ( ) in place of $ R (r , ) .


5.3 New Measurement Techniques

In this section a new approach is proposed to interpret surface wave measurements.


The first and most important justification for the new approach is that it provides
consistency between phase velocity and attenuation measurements. As shown in Section 5.2
the conventional procedures used to determine the experimental dispersion and attenuation
curves are quite different. The dispersion curve is obtained from the repeated application of
the two-station method in which the results of several different receiver spacings are averaged
to obtain a single composite dispersion curve. Conversely, the experimental attenuation
curve is determined from a procedure based on the application of a multi-station method,
where the individual attenuation coefficients are calculated from a non-linear regression
involving measured displacement amplitudes. One of the objectives of the next two sections
is to illustrate a procedure in which the experimental dispersion and attenuation curves are
both determined with the multi-station method.
The second reason motivating a new interpretation approach comes from the
formulation of the coupled inversion presented in Section 4.7. As mentioned at the
beginning of this chapter, in a truly consistent approach to surface wave testing, the
dispersion and attenuation curves should be both measured and inverted simultaneously.
The objective of Section 5.3.2 is to illustrate a technique for the simultaneous determination
of the dispersion and attenuation curves from surface wave measurements.
5.3.1

Uncoupled Measurements

From Eq. 3.76 the vertical displacement w(r, ) induced in a linear viscoelastic
vertically heterogeneous medium by a harmonic source Fy e it located at the ground
surface is given by:
w(r , ) = Fy G (r , ) e [

i t * ( r , )

(5.7)

Rayleigh Phase Velocity and Attenuation Measurements

139

where w(r , ) = u y (r ,0 , ) , G (r , ) = Gyv (r ,0 , ) , and * (r , ) = yv* (r ,0 , ) . In Section

3.7 it was shown that the term * (r , )

,r

can be interpreted as an effective complex

$ R ( r , ) . By decomposing
=
+
i

,r
V$ R ( r , )

* (r , ) into its real and imaginary parts, Eq. 5.7 can be rewritten as follows:

[ (r , )]
*

wavenumber and written as

w(r , ) = Fy G (r , ) e 2 ( r , ) e [

i t 1 ( r , )

(5.8)

where 1 (r , ) = * (r , ) , and 2 (r , ) = * (r , ) . Considering only the spectral


part of Eq. 5.8 gives:
w(r , ) = Fy G (r , ) e 2 ( r , )

arg w(r , ) = 1 (r , )

(5.9)

Since V$ R (r , ) =

1 (r , )

and $ R ( r , ) = 2 (r , )

,r

from Eq. 5.9 it is possible to

,r

obtain:

$
VR (r, ) =
arg w(r , )

w(r , )
$ (r , ) = ln

R
r G (r , )

( [

])

(5.10)

Equation 5.10 suggests an interesting geometrical interpretation. If the angular frequency


is constant, the effective Rayleigh phase velocity V$ R (r , ) is proportional to the inverse
of the slope of the displacement phase plotted as a function of the source-to-receiver
distance r. Moreover Eq. 5.10 also shows that at constant frequency, the effective Rayleigh
attenuation coefficient $ R (r , ) can be interpreted as the slope (with changed sign) of a

140

Rayleigh Phase Velocity and Attenuation Measurements

corrected displacement amplitude, the latter considered again as a function of the source-toreceiver distance r. The corrected displacement amplitude is defined by the natural

w(r , ) G (r , ) . Figure 5.5 illustrates this important geometrical


interpretation of the effective Rayleigh phase velocity V$ R (r , ) and attenuation coefficient

logarithm of

arg[w(r, )]

$ R (r , ) . Equation 5.10 is important because it suggests a procedure for the experimental


determination of V$ R (r , ) and $ R (r , ) .

= constant
V$ R (r* , ) =

k$ R (r * , )
r*

r
k$ R

Figure 5.5(a)

Geometrical Interpretation of Effective Rayleigh Phase Velocity

This procedure requires the measurement of displacement phase and amplitudes in a


linear receiver array (which are given by V (r , ) [ C( )] ), for a specified set of
frequencies. The measured displacement phase is then plotted at each frequency versus the r
at each frequency. If the number of receivers is sufficiently large to define the curve well,
Fig. 5.5(a) shows how to determine the effective Rayleigh phase velocity V$ R (r , ) .
A plot of corrected displacement amplitude versus source-to-receiver distance can be
used to determine the effective Rayleigh attenuation coefficient $ R (r , ) as shown in Fig.
5.5(b).

Rayleigh Phase Velocity and Attenuation Measurements

ln[|w(r,)|/G(r,) ]

= constant
$ R = $ R (r* , )
1

$ R

r*

Figure 5.5(b)

141

Geometrical Interpretation of Effective Rayleigh Attenuation Coefficient

In the latter case however, it is assumed that the V$ R (r , ) data have already been

inverted so that the geometric spreading function G (r, ) required to calculate the
corrected displacement amplitudes is known (for weakly dissipative media).
Although this procedure is in principle correct, its practical implementation may be very
difficult due to the unstable numerical differentiation required by Eq. 5.10. This problem
associated with the experimental determination of the dispersion and attenuation surfaces
has already been discussed in Section 4.6.4, where a possible solution strategy has also been
presented.
Equation 5.10 demonstrates that the effective Rayleigh phase velocity and attenuation
coefficient are indeed derived quantities that are obtained from differentiation of the
experimentally measured displacement phases and amplitudes. This consideration is the
basis of a new interpretation of the SASW test where the inverted quantities are
displacement phases and amplitudes rather than phase velocities and attenuation
coefficients. The details of this new interpretation were illustrated in Sections 4.6.4 and
4.7.4. If in Eq. 5.7 the complex phase angle * (r , ) is approximated by

* (r , ) K * ( ) r with K * ( ) = k R ( ) + i R ( ) , Eq. 5.9 becomes:

142

Rayleigh Phase Velocity and Attenuation Measurements


w(r , ) = Fy G (r , ) e R ( )r

arg w(r , ) = k R ( ) r

(5.11)

Equation 5.11 suggests a method for estimating the average effective Rayleigh phase
velocity and attenuation coefficient over the same linear array of receivers. The method
simply consists of computing k R ( ) and hence VR ( ) from a linear regression involving
the experimental displacement phases and then inverting VR ( ) to obtain the shear wave
velocity profile and G (r, ) . The attenuation coefficient R ( ) is obtained from a nonlinear regression involving the displacement amplitudes. The procedure for determining
R ( ) is the same as the conventional method presented in Section 5.2.2. The key
difference is that the same configuration of receivers is used to obtain both VR ( ) and
R ( ) , and therefore the method suggested by Eq. 5.11 provides consistency between
surface wave phase velocity and attenuation measurements.
5.3.2

Coupled Measurements

In this section an alternate approach to velocity and attenuation measurements in SASW


testing is presented. This approach is based on the concept of a displacement transfer function
that allows the simultaneous determination of both Rayleigh phase velocity and the
attenuation coefficient.
In a linear system, which in this case corresponds to a linear viscoelastic soil deposit, the
ratio between an output and an input signal in the frequency domain is called the frequency
response function or the transfer function of the system (Oppenheim and Willsky, 1997). In
the typical SASW test configuration shown in Fig.5.1, the input signal is the harmonic force
applied by the vertically oscillating source Fy e it , while the output signal is the vertical
displacement w(r , ) measured at a distance r from the source.
From Eq. 5.7, the displacement transfer function T (r, ) between source and receiver
is given by:

T (r , ) =

w( r , )
i * ( r , )
it = G ( r , ) e
Fy e

(5.12)

Because the dynamic signal analyzer allows direct measurement of T (r, ) , Eq. 5.12 can
be used as a basis of a non-linear regression analysis to determine the complex phase angle

Rayleigh Phase Velocity and Attenuation Measurements

143

* (r , ) from the experimentally measured displacement transfer functions T (r, ) . From


the knowledge of the phase angle * (r , ) the effective Rayleigh phase velocity V$ (r , )
and

attenuation

[ (r , )]
*

,r

coefficient

$ R (r , )

may

be

computed

from

the

relation

=
+ i $ R ( r , ) . Note that in this procedure V$ R (r , )
V$ R ( r , )

and

$ R (r , ) are determined simultaneously.

As in the case of the uncoupled measurements, this procedure, although in principle


correct, is difficult to implement from a practical point of view because of the need to
perform unstable operations of numerical differentiation. Furthermore, there is an
additional difficulty in simultaneously measuring V$ R (r , ) and $ R (r , ) because the
geometric spreading function G (r, ) in Eq. 5.12 is unknown.

In vertically heterogeneous, linear viscoelastic media G (r, ) is a complicated function


of the medium parameters that the SASW test seeks to determine. The problem can be
overcome by an iterative strategy that combines Eq. 5.12 with an inversion algorithm, and
which starts from a tentative profile of material parameters and proceeds until convergence.
Since in most cases the magnitude of G (r, ) is controlled by a term proportional to 1 r
(which is the Rayleigh geometric spreading law in homogeneous media), the iterative
scheme requires relatively few iterations to converge.
The simultaneous measurement of dispersion and attenuation data may be considerably
simplified by assuming * (r , ) K * ( ) r in Eq. 5.12. In this case the experimental
displacement transfer functions
T ( r , ) = G ( r , ) e i K

( ) r

T (r, )

could be inverted using the equation

to yield K * ( ) =
+ i R ( ) directly.
VR ( )

5.4 Statistical Considerations


5.4.1

Overview

Uncertainty in surface wave measurements is an important consideration. In


conventional SASW testing the measured quantities are the auto-power spectrum G rr ( )
of each receiver location and the cross-power spectrum between a pair of receivers
G r1r2 ( ) . In the new coupled measurement technique presented in the last section, the
measured quantity is the particle velocity transfer function H(r, ) between source and

144

Rayleigh Phase Velocity and Attenuation Measurements

receiver, which is defined by H(r , ) = w


& (r , ) Fy e it = iT(r , ) . These experimental
quantities are assumed to be normally distributed.
The validity of this assumption may be arguable, but the uncertainties associated with
the methods used to interpret surface wave measurement do not justify the recourse to
more refined statistical distributions in the opinion of the writer. The statistics of normally
distributed experimental data are the expected value and variance. Bias and other systematic
errors in the measurements are not considered.
The recording device used in surface wave testing (generally a dynamic signal analyzer)
determines an estimate of the expected values of G rr ( ) , G r1r2 ( ) , and H(r, ) , which are

indicated by E G rr ( ) , E G r1r2 ( ) , and E H(r, ) . It is also possible to compute the

variances of these measured quantities using the following relations (Bendat and Piersol,
1986):

G 2rr ( )
(
)
Var[G rr ] =
nd

(
)
G

r
r

12
Var G r1r2 ( ) = n 2 ;
d
r1 r2

1 sr2 ) H(r , )
(
Var H(r , )
2 n d sr2

(1 sr2 )

Var arg H(r , )


2 sr2

[ {

}]

[ {

}]

Var arg G r1r2 ( )

(1 )

2
r1r2

(5.13a)

2 r21r2

(5.13b)

where sr ( ) represents the ordinary coherence function between the harmonic source
and the receiver output signal. The number n d is the number of independent averages used
to estimate the spectral quantities.
2

2
2
The variance of the ordinary coherence function r1r2 ( ) or sr ( ) can be estimated
by the following relation (Bendat and Piersol, 1986):

Rayleigh Phase Velocity and Attenuation Measurements

[ ]

Var r21r2

5.4.2

2 r21r2 1 r21r2

145

(5.14)

nd

Statistical Aspects of Conventional Measurements

Once expected values and variances of the experimental measurements have been
defined, the tools of statistical analysis can be used to estimate the expected values and
variances of the quantities depending on them. In conventional phase velocity
measurements the first of these quantities is the Rayleigh phase velocity VR ( ) that is
obtained from the phase of the cross-power spectrum Gr1r2 via Eq. 5.2. Because VR ( ) is

( )

a non-linear function of arg Gr1r2 , VR ( ) is in general non-Gaussian distributed. However,


by assuming the experimental data to be characterized by small variances, it is possible to

[ ( )] (provided E[arg(G )] 0 ), and

expand Eq. 5.2 in a Taylors series about E arg G r1r2

r1r2

truncate it to first-order terms only (Papoulis, 1965). If it is further assumed that

Var[ ] = Var r2 r1 = 0 (i.e. frequency of excitation and receiver spacing are considered
deterministic rather than random variables), the following results are obtained:

(r2 r1 )
E VR ( )

E arg G r1r2 ( )

Var arg G r1r2 ( ) (r2 r1 )

Var VR ( )
4

(
)

E
arg
G
r1r2

[ (

)]

]
[ ( )] [
{ [ ( )]}

(5.15)
2

Since the experimental dispersion curve is constructed by averaging Rayleigh phase

velocities VR ( )

] (j = 1, n )
j

obtained over n R different receiver spacings, and since

every linear combinations of normal distributions is itself a normal distribution (Papoulis,


1965), the expected value and variance of the experimental dispersion curve VR ( ) are
given by:

146

Rayleigh Phase Velocity and Attenuation Measurements

1 nR
(
)

=
E
V
E VR ( ) j

R
n R j =1

nR
Var V ( ) = 1
Var VR ( )
R

n R2 j=1

(5.16)
j

In Eq. 5.16 the variance of VR ( ) has been computed under the assumption that the

individual VR ( ) j are statistically uncorrelated.


In conventional attenuation measurements, the displacement amplitudes w (r, ) are
calculated from the experimental auto-power spectra via Eq. 5.5. Using the same procedure
illustrated above for VR ( ) , namely expanding Eq.(5.5) in a truncated Taylors series about

[ ]

E G rr and assuming Var[C( ) ] = 0 , it is possible to obtain:

E G rr ( )
E w(r , )

C( )

Var G rr ( )

Var w(r , )
4 E G rr ( ) C( )

][

(5.17)
2

The expected value and variance of the Rayleigh attenuation coefficient R ( ) must be
computed from the non-linear regression based on Eq. 5.6. Unfortunately, because of the
non-linearity of the relationship between w(r, ) and R ( ) , one can only estimate

E R ( ) and Var R ( ) . If the attenuation coefficient is estimated from Eq. 5.6 using
a standard non-linear least-squares algorithm (e.g. the Gauss-Newton method or the
Levenberg-Marquardt method), the uncertainty associated with this estimate can be
approximately calculated with the following relation (Menke, 1989):

] (

Var R ( ) J T R J R

](

J TR Cov w ( ) J TR J R
last #

J TR
last #

(5.18)

where Cov [ w ( ) ] is an n T n T matrix containing the covariances of the experimental


displacement amplitudes at the n T different receiver locations at a given frequency .

Rayleigh Phase Velocity and Attenuation Measurements

147

Since it is assumed that the data w (r , ) are uncorrelated, the data covariance matrix is

diagonal with the non-zero elements equal to the variances of w (r , ) given by Eq. 5.17.

( ) (k = 1, n ) are defined from

The term J R is a n T 1 vector whose components J R

Eq. 5.6 by:

(J ) = [J
R

() =

w ( rk , )
R

= rk w ( rk , )

(5.19)

The subscript last # outside the brackets of Eq. 5.18 indicates that the terms inside the
parentheses, essentially the vector J R , refer to the last iteration in the solution of the nonlinear regression based on Eq. 5.6.
Equation 5.18 completes the calculation of the expected values and variances of the
experimental Rayleigh phase velocities VR ( ) and attenuation coefficients R ( ) . The
next task is to determine how the uncertainties of these data are mapped into uncertainties of
the estimated model parameters, which are the shear wave velocity and shear damping ratio
profiles of a given soil deposit. This and other related topics will be discussed in Sections
5.4.4 and 5.4.5.
5.4.3

Statistical Aspects of New Measurement Techniques

5.4.3.1 Uncoupled Analysis


In the new approach to surface wave measurements, the average Rayleigh phase velocity
VR ( ) in the simplified uncoupled analysis is computed from a linear regression of the
displacement phases versus receiver distance at constant frequency. More precisely, the
method consists of determining the Rayleigh wavenumber k R ( ) from a linear regression

based on Eq. 5.11b, and then obtaining VR ( ) = k R ( ) . By estimating E k R ( ) using


a standard linear least-squares algorithm, it is a straightforward matter to calculate the
uncertainty associated with this estimate using the following relation (Menke, 1989):

] [(G

Cov m ( ) = (G rT G r ) G rT Cov arg ( w ( ) )


1

T
r

G r ) G rT
1

(5.20)

148

Rayleigh Phase Velocity and Attenuation Measurements

where Cov arg ( w ( ) ) is an n T n T diagonal matrix (assuming the data arg w (r , ) to


be uncorrelated) whose elements are the variances of the experimental displacement phases

arg w (r , ) at the n T different receiver locations at a given frequency .

The term Cov m ( ) is a 2 2 diagonal matrix whose two non-zero elements are the

variances of the intercept and the slope (i.e., Var k R ( ) ) of the linear regression. Finally,
T

r1 r2 L rn T
the term G r is a n T 2 matrix defined by G r =
where rk ( k = 1, n T )
1 1 L 1
denotes the source-to-receiver distance for receiver k.

Having defined E k R ( ) and Var k R ( ) , the expected value and variance of


VR ( ) = k R ( ) are determined using the Taylors series expansion described in the
previous section. The results are given by the following expression:

( )
E VR E k ( )
R

2 Var k R ( )

4
Var VR ( )
E k R ( )

{[

]}

(5.21)

Expected values and variances of the Rayleigh attenuation coefficient R ( ) are


determined with the same procedure used for the conventional measurements; in particular
Eq. 5.18 gives a measure of the uncertainty associated with the estimate of R ( ) .
5.4.3.2 Coupled Analysis
In coupled measurements of dispersion and attenuation data, the quantity measured
experimentally is the particle velocity transfer function H(r, ) between source and

receiver, and defined by H(r , ) = w


& (r , ) Fy e it = iT(r , ) . The expected value and

variance of the modulus and phase of H(r, ) are defined by Eq. 5.13b. The corresponding
statistical quantities for the displacement transfer function T (r, ) are:

Rayleigh Phase Velocity and Attenuation Measurements

[ (

E T (r , ) = E H(r , ) ;

2
Var T(r , ) = Var H(r , ) ;

149

) ] = E [arg (H(r , ))] + 2

E arg T (r , )

[ (

) ] = Var [arg (H(r , ))]

(5.22)

Var arg T (r , )

Now, since the complex wavenumber K * ( ) =


+ i R ( ) is determined from
VR ( )

the non-linear regression T (r , ) = G (r , ) e iK

[
[

] [
] [

( ) r

, it is first necessary to calculate

E T(r, ) and Var T(r, ) in order to compute expected value and variance of K * ( ) .

] { [ (
] { [ (

) ]}
) ]}

It is convenient to write T (r , ) = T1 (r , ) + iT2 (r , ) , from which it is simple to obtain:


E T (r , ) E T (r , ) cos E arg T(r , )
1

E T2 (r , ) E T(r , ) sin E arg T(r , )

Var T(r , ) = Var T1 (r , ) Var T2 (r , )

(5.23)

where:

] cos [arg (T(r , ))]Var [ T(r , ) ] +


+ T(r , ) sin [arg (T (r , ))]Var[arg ( T(r , ))]

Var T (r , )
1

Var T2 (r , )

] sin [arg(T(r , ))]Var [ T(r , ) ] +


+ T (r , ) cos [arg ( T(r , ))]Var[arg ( T(r , ))]

(5.24)

The uncertainty of the complex wavenumber K * ( ) is finally computed from the non-

linear regression T (r , ) = G (r , ) e iK ( )r following a procedure that is formally identical


to that used for determining the uncertainty of the Rayleigh attenuation coefficient R ( ) .
The result is:
*

150

Rayleigh Phase Velocity and Attenuation Measurements

] (

Var ( ) J HK * J K *

](

J Cov T( ) J HK * J K *
last #

H
K*

J
last #
H
K*

(5.25)

where Cov [ T( ) ] is an n T n T matrix representing the covariances, at a given frequency


, of the experimental displacement transfer functions at the n T receiver locations. Since

it is assumed that the data T (r, ) are uncorrelated, the matrix Cov [ T( ) ] is diagonal
with the non-zero elements equal to the variances of T (r, ) and given by Eqs. 5.23 and
5.24. The term J K * is an n T 1 complex-valued vector whose components

(j = 1, n ) are defined from the relation T(r , ) = G (r , ) e


T

(J ) = [J
K

() =

iK ( ) r
*

( ) = i r T r ,
( )
K
j

K*

and are equal to:

T rj ,

(J )

(5.26)

Like Eq. 5.18, the subscript last # outside the brackets of Eq. 5.25 indicates that the
terms inside the parenthesis refer to the last iteration in the solution of the non-linear

regression T (r , ) = G (r , ) e iK

( ) r

Once E [ K * ( ) ] and Var [ K * ( ) ] are computed, the expected value and variance of

the complex Rayleigh phase velocity VR* ( ) are obtained from the complex extension of
Eq. 5.21, namely:

*
( )
E VR E K * ( )
R

2 Var K *R ( )

*
4
Var VR ( )
E K *R ( )

{[

]}

(5.27)

Equation 5.27 completes the statistical analysis of surface wave measurements when
dispersion and attenuation data are determined simultaneously from the particle velocity

transfer functions H(r, ) . However it may also be important to compute E VR* ( ) and

Rayleigh Phase Velocity and Attenuation Measurements

151

Var VR* ( ) when surface wave data, namely VR ( ) and R ( ) , are obtained
independently. The solution to this problem is obtained from the application of the theory
of random variables (Papoulis, 1965) to Eq. 2.34.
In particular, this equation can be interpreted as a mapping : R 2 C that assigns a
complex random variable VR* to a pair of independent random variables VR and D R . If

the variables VR and D R are normally distributed and the mapping (VR , D R ) is
sufficiently smooth, (VR , D R ) may be expanded in a Taylor series about the point

[E(V ), E(D )] . Again assuming small variances for the variables V


R

and D R , this series

is truncated to first-order terms only, yielding:

] (
]

])

[
[

]
]

E VR ( )
E VR* ( )
1 iE D R ( )
1 + E 2 D R ()

1 iE D R ( )
*
Var VR ( )
Var VR ( ) +
2

1 + E D R ( )

E V ( ) 2E D ( ) + i 1 E 2 D ( )
R
R
R

+
2

1 + E 2 D R ( )

][ [

] (

])

(5.28)

])]

Var D R ( )

where:

E R ( ) E
E D R ( )

E VR ( )

Var D R ( )

] [V

( )]
(5.29)

E R ( )
Var R ( ) +

Var VR ( )

In deriving Eqs. 5.28 and 5.29 it was assumed that the random variables VR and D R are
uncorrelated.

152

Rayleigh Phase Velocity and Attenuation Measurements

In summary, despite the fact that the relationship between directly measured quantities
(i.e., G rr ( ) , G r1r2 ( ) , and H(r, ) ) and the derived surface wave data (i.e., VR ( ) ,

R ( ) , and VR* ( ) ) is non-linear in most cases, the assumption of small variances allows
one to obtain explicit results for the expected values and variances of VR ( ) , R ( ) , and
VR* ( ) .
5.4.4

Statistical Aspects of Uncoupled Rayleigh Inversion

Having defined the statistics of the experimental data, the next task is to determine how
the uncertainty of the measurements is mapped into uncertainty of the estimated model
parameters, which are the shear wave velocity VS and shear damping ratio D S profiles. The
first objective will be to obtain an approximation of the uncertainty associated with the
inverted shear wave velocity profile VS . The problem of inverting an experimental

dispersion curve can formally be written as VR = VR ( VS ) , where VR is a n F 1 vector of


Rayleigh phase velocities VR ( ) associated with n F different frequencies, and VS is a
n L 1 vector whose components are the unknown shear wave velocities of an n L layers in
the soil deposit.

[ ] and Var [ V ] are the quantities that are of


= V ( V ) is non-linear, only approximate results can

From a statistical point of view, E VS


interest. Because the relationship VR

be obtained. If VR = VR ( VS ) is inverted using Occams algorithm, it is possible to show

[ ]

that the uncertainty associated with the estimated E VS


following relationship:

Cov VS T + WVR J S

[ ]

WVR J S

Cov VR T + WVR J S

[ ]

(W J )
VR

WVR J S

can be computed with the

WVR
last #

(W J )
VR

WVR
last #

(5.30)

[ ]

where Cov VR is an n F n F matrix of covariances of the experimental Rayleigh phase


velocities at n F different frequencies . It is assumed that the data VR ( ) are

[ ]

uncorrelated, and hence the matrix Cov VR is diagonal with the non-zero elements equal
to the variances of VR ( ) . The variances are given by Eq. 5.16 for conventional
measurements and by Eq. 5.21 for the new measurement procedure.

Rayleigh Phase Velocity and Attenuation Measurements

[ ]

Cov VS

153

is a n L n L diagonal matrix whose elements are the variances of the

estimated shear wave velocities ( VS ) j with j = 1, n L . The term WVR is a n F n F diagonal


matrix

( )
VR

defined

by

( ) , 1 / ( ) ,..... ,1 / ( )

WVR = diag 1 / VR

VR

VR

nF

where

= + Var ( VR ) k with k = 1, n F and represents the standard deviations associated with

the experimental data VR . The parameter and the symbol have already been defined
in Section 4.5; is the inverse of the Lagrange multiplier and is the two-point-central
finite difference operator.
Finally, the term J S is the n F n L Jacobian matrix whose elements ( JS ) kj are defined

by the partial derivatives ( JS ) kj = ( VR ) k ( VS ) j . The subscript last # outside the brackets


of Eq. 5.30 indicates that the terms inside the parentheses (in essence the Jacobian matrix
J S ) should be computed with respect to the last iteration in the solution of the non-linear
problem VR = VR ( VS ) .

The determination of the uncertainty associated with the estimated damping ratio

[ ]

profile D S is simpler than determining Cov VS because the inversion of the experimental
attenuation curve R ( ) to obtain the damping ratio profile is linear. The problem can
symbolically be written as GD S = R . In this equation G is a n F n L matrix formed by
the partial derivatives of Rayleigh phase velocity with respect to the shear and compression
wave velocities of the soil layers and defined by Eqs. 3.45 and 3.63b. The term R is a
n F 1 vector of Rayleigh attenuation coefficients calculated for n F different frequencies
and D S is a n L 1 vector containing the unknown shear damping ratios of the n L layers
soil deposit. If the linear equation GD S = R is inverted using Occams algorithm, the
uncertainty of the estimated shear damping ratio profile D S can be computed from the
uncertainty of the measured attenuation coefficients R using the following relation:

1
T
T

Cov D S = T + W R G W R G W R G W R

[ ]

[ ]

Cov R

1
T
T
T

+ W R G W R G W R G W R

(5.31)

154

Rayleigh Phase Velocity and Attenuation Measurements

[ ]

where Cov R is an n F n F diagonal matrix whose non-zero elements (the R ( ) are


assumed to be uncorrelated) are the variances of R ( ) at n F different frequencies ; the

[ ]

variances can be calculated from Eq. 5.18. Cov DS

is a n L n L diagonal matrix

containing the variances of the estimated shear damping ratios (DS ) j with j = 1, n L .
Finally,

the

term

( )

is

W R

nF nF

( )

diagonal

( )

matrix

defined

by

( )

W R = diag 1 / R , 1 / R ,..... ,1 / R where R = + Var ( R ) k with


1
2
nF
k

k = 1, n F and represents the standard deviations of the experimental data R .


5.4.5

Statistical Aspects of Coupled Rayleigh Inversion

The procedure used to calculate an approximation of the uncertainty associated with the
inverted complex shear wave velocity profile VS* is identical to that described in the

[ ]

previous section to obtain Cov VS . The only difference is that the computation of

[ ] requires a systematic use of the complex formalism.


*
S

Cov V

The non-linear relation between a complex-valued vector VS* of n L unknown shear


wave velocities (all referred to a reference frequency ref ), and a n F 1 vector of complex

Rayleigh phase velocities VR* for n F different frequencies is denoted by VR* = VR* ( VS* ) . As
for the real case, the non-linearity of this relationship allows one to obtain only approximate

[ ]

results for Cov VS* . If the equation VR* = VR* ( VS* ) is inverted using Occams algorithm,
the uncertainty of the estimated VS* profile can be computed from the complex version of
Eq. 5.30, namely:

Cov V T + WV* * J *S
R

[ ]
*
S

W J

*
VR*

Cov VR* T + WV* * J*S


R

[ ]

[ ]

where Cov VR*

*
S

(W J ) W
*
VR*

)W
H

*
VR*

*
S

J *S

*
VR*

last #

(W J )
*
VR*

*
S

WV* *
R
last #

(5.32)

is an n F n F diagonal matrix (for uncorrelated VR* ( ) data) whose

elements are the (complex) variances of the experimental VR* ( ) at n F different

Rayleigh Phase Velocity and Attenuation Measurements

155

[ ]

frequencies and given by Eq. 5.27. Cov VS* is a n L n L diagonal matrix formed by the

variances of the estimated complex valued shear wave velocities ( VS* ) j ( j = 1, n L ) .


The

term

WV* *

is

( ) ( )

nF nF

( )

diagonal

matrix

defined

by

( )

WV* * = diag 1 / *V * , 1 / *V * ,..... , 1 / *V * where *V * = + Var ( VR* ) k with


R
R 1
R
R
R
nF
2
k

k = 1, n F and represents the complex-valued standard deviations of the experimental VR* .

Finally, the term J *S is the n F n L complex-valued Jacobian matrix whose elements ( J*S ) kj
are defined by the partial derivatives ( J*S ) kj = ( VR* ) k ( VS* ) j .

156

Rayleigh Phase Velocity and Attenuation Measurements

6 VALIDATION OF THE ALGORITHMS

6.1 Overview
The objective of this chapter is the validation of the algorithms developed during this
research study, and based on some of the theoretical concepts presented Chapters 3 and 4.
The validation of these algorithms is carried out using the following procedure. For a
layered medium characterized by a set of material parameters, synthetic displacement data at
several receiver offsets are generated from the numerical solution of the Rayleigh forward
problem. These surface wave displacement data, which simulate SASW field measurements,
are then used to calculate synthetic dispersion and attenuation curves with the same
techniques used in a real SASW test and illustrated in chapter 5.
The simulated dispersion and attenuation curves are then inverted using the Rayleigh
inversion algorithms described in Chapter 4 in an attempt to recover the original shear wave
velocity and shear damping ratio profiles of the medium. A comparison between the
original and the inverted VS and D S profiles will provide the means for assessing the
performance of the algorithms. This validation procedure is carried out for stratified media
having different layering and material properties.
This chapter is organized in two main sections, describing the procedures adopted for
the validation of the uncoupled and coupled inversion algorithms, respectively. In each of these
sections, the algorithms associated with the fundamental mode analysis (i.e. UFUMA and
CFUMA) are presented first, followed by those based on the equivalent multi-mode analysis (i.e.
UEQMA and CEQMA).
6.2 Lambs Problem
Before proceeding with the analyses for layered soil deposits, a few results will be
presented for the solution of the Rayleigh boundary value problem in homogeneous elastic
and viscoelastic media. For this simple case, an exact solution is available which can be used
to verify the results obtained from the algorithms developed for the solution of the elastic
and viscoelastic Rayleigh forward problem.
The problem of determining the displacement field induced by a vertical harmonic point
load applied at the free surface of an homogeneous, isotropic, linear elastic half-space, was
first solved by Lamb (1904) in a classical paper entitled On the Propagation of Tremors over the
Surface of an Elastic Solid. Lamb used the tools of complex variable theory to find the
solution of what today is known as the Lambs problem, which can be considered as the
dynamic analogue of another classical problem of linear elasticity: Boussinesqs problem.

157

158

Validation of the Algorithms

Although Lamb developed a solution for an arbitrary time variation of the source, only
the solution for a harmonic source is presented here. The harmonic solution can be written
as follows:
Fe it
w R (r , ) =
k (k R ) H (02 ) ( k R r )
2 iG R

(6.1)

where w R (r , ) is the Rayleigh vertical displacement at the free surface of a homogeneous


elastic half-space at a distance r from a vertically oscillating harmonic source Fe it . The term
G is the shear modulus of the elastic medium, k R is the Rayleigh wave-number, and the
symbol H (02 ) () denotes the Hankel function of the second kind of zero order (i.e.
H (02 ) ( z) = J0 (z ) iY0 ( z) where J0 (z ) and Y0 ( z) are the Bessel functions of the first kind
and second kind, respectively, of zero order).
The function (k R ) is defined as follows:

( k R ) =

k S2 k R2 k P2

(6.2)

R ( k R )

where k P and k S are the wave-numbers of the P-wave and S-wave, respectively. Finally, the

function R ( k R ) is given by the following expression:


R ( k R ) = (2 k 2R k 2S ) 4 k 2R
2

(k

2
R

k 2P )( k 2R k 2S )

(6.3)

where k R = VR and VR is the frequency-independent Rayleigh phase velocity to be


determined from the solution of the Rayleigh dispersion equation in homogeneous media,
namely:
3 8 2 + 8 (1 + 2 ) 16 = 0

(6.4)

where = (VR VS ) , = 1 ( VS VP ) , and V = k with = P , S .


2

Validation of the Algorithms

159

Lambs solution was calculated for three different elastic media, and the results were
compared with those obtained with the new algorithm. Table 6.1 shows the material
properties and frequencies used for the comparison test.
Table 6.1

Medium Properties and Frequencies Used for Validation of the Elastic


Lambs Problem

Mass
Density

VP

VS

[m/s]

[m/s]

500

250

1.8

600

400

1.8

50

1000

600

1.8

100

Case No.

[t/m3]

Frequency
[Hz]

The results of the validation test are shown in Fig. 6.1(a) through Fig. 6.1(c), where
vertical displacement amplitudes and phases are plotted versus the distance from the source
for a specified frequency.
x 10

-7

6
Lamb Solution
Numerical Method

0.8

Lamb Solution
Numerical Method

Displacement Phase [rad]

Displacement Amplitude [m]

0.6

0.4

0.2

0
0

20

Figure 6.1(a)

40
60
Distance [m]

80

100

2
0
-2
-4
-6
0

20

40
60
Distance [m]

80

100

Comparison of Solutions for the Elastic Lambs Problem (Case 1)

The magnitude of the harmonic source Fe it has been assumed equal to one. In all
three cases the agreement between the closed-form solution given by Lamb and that
obtained with the numerical algorithm is excellent.

160

Validation of the Algorithms

x 10

-7

0.6

0.4

0.2

20

Figure 6.1(b)

Displacement Amplitude [m]

Displacement Phase [rad]

Lamb Solution
Numerical Method

x 10

40
60
Distance [m]

80

0
-2
-4

20

40
60
Distance [m]

80

100

Comparison of Solutions for the Elastic Lambs Problem (Case 2)

-7

0.8

Lamb Solution
Numerical Method

0.6

0.4

0.2

0
0

-6
0

100

20

Figure 6.1(c)

40
60
Distance [m]

80

100

Lamb Solution
Numerical Method

4
Displacement Phase [rad]

Displacement Amplitude [m]

0.8

Lamb Solution
Numerical Method

2
0
-2
-4
-6
0

20

40
60
Distance [m]

80

100

Comparison of Solutions for the Elastic Lambs Problem (Case 3)

The solution of the Lambs problem in viscoelastic media can be obtained from the
application of the elastic-viscoelastic correspondence principle to the corresponding elastic
solution. As a result, in the viscoelastic solution the terms k P , k S , k R and G appearing in
Eq. 6.1 are complex-valued. Table 6.2 shows the frequencies and material parameters used
for the validation of the numerical solution of the Lambs problem in viscoelastic media.
Figures 6.2(a) through 6.2(c) show the results of the comparison. It is apparent from the
plots that the closed-form and numerical solutions also agree for the viscoelastic case.

Validation of the Algorithms


Table 6.2

Medium Properties and Frequencies Used for Validation of the Viscoelastic


Lambs Problem

VP

VS

DP

DS

Mass
Density

[m/s]

[m/s]

[%]

[%]

[t/m3]

500

250

1.8

600

400

1.8

50

1000

600

1.8

100

Case No.

x 10

Displacement Amplitude [m]

Lamb Solution
Numerical Method

0.6

0.4

0.2

20

Figure 6.2(a)

40
60
Distance [m]

80

2
0
-2
-4
-6
0

0
100

20

80

100

Comparison of Solutions for the Viscoelastic Lambs Problem (Case 1)


6
Lamb Solution
Numerical Method

4
Displacement Phase [rad]

Lamb Solution
Numerical Method

0.6

0.4

0.2

0
0

40
60
Distance [m]

-7

0.8
Displacement Amplitude [m]

Lamb Solution
Numerical Method

x 10

[Hz]

0.8

Frequency

-7

Displacement Phase [rad]

161

20

Figure 6.2(b)

40
60
Distance [m]

80

100

2
0
-2
-4
-6
0

20

40
60
Distance [m]

80

100

Comparison of Solutions for the Viscoelastic Lambs Problem (Case 2)

162
x 10

-7

Displacement Amplitude [m]

0.8

Lamb Solution
Numerical Method

0.6

0.4

0.2

Lamb Solution
Numerical Method

20

Figure 6.2(c)

40
60
Distance [m]

80

100

Displacement Phase [rad]

Validation of the Algorithms

2
0
-2
-4
-6
0

20

40
60
Distance [m]

80

100

Comparison of Solutions for the Viscoelastic Lambs Problem (Case 3)

6.3 Numerical Simulations


For vertically layered media the validation of the inversion algorithms was conducted
according to the following procedure. Three soil profiles were selected as simplified
examples of actual soil deposits. Tables 6.3 through 6.5 show the material properties and
layer thicknesses that were chosen for the three soil profiles. Case 1 represents a regular soil
profile where the stiffness of the layers increases with depth. Cases 2 and 3 represent two
irregular soil profiles. Case 2 represents the common situation in which a soft layer is
trapped between two stiffer layers. Finally, Case 3 is an example of a medium characterized
by a stiff surface layer. The soil profile of Case 1 belongs to the category of normally dispersive
soil profiles (Tokimatsu, 1995), while Cases 2 and 3 are classified as inversely dispersive soil
profiles.
As discussed in Section 3.4, near-field effects decay very quickly with distance from the
source in normally dispersive media (they are negligible at distances larger than one-half a
wavelength). In inversely dispersive media, such effects are considerably more important
(they may extend up to distances that are four times larger). Furthermore, in normally
dispersive media the propagation of surface waves is to a great extent dominated by the
fundamental mode, while in inversely dispersive soil deposits the response of the medium is
more often controlled by higher modes of propagation, particularly at high frequencies
(Gucunski and Woods, 1991). For these three layered media, a numerical simulation was
conducted to compute synthetic dispersion and attenuation curves.
This computation was performed using the computer programs developed for the
solution of the Rayleigh forward problem and the methods for surface wave data
interpretation described in Section 5.3. The synthetic displacement field was computed at 20
receiver locations, with the closest receiver located at a distance equal to one to two
wavelengths from the source. The numerical simulation was performed for a set of 50

Validation of the Algorithms

163

frequencies ranging from 5 to 100 Hz, which is a typical frequency range used in SASW
tests. Once the synthetic dispersion and attenuation curves of the three soil deposits were
determined, the last step of the validation procedure consisted of inverting these curves to
obtain the corresponding shear wave velocity and shear damping ratio profiles. A
comparison of the latter with the original VS and D S profiles served as the basis for
assessing the performance of the inversion algorithms.
Table 6.3

Medium Properties Used for the Validation of the Inversion Algorithms


(Case 1)

VP

VS

DP

DS

Mass
Density

[m/s]

[m/s]

[%]

[%]

[t/m3]

400

200

2.0

3.5

1.7

10

600

300

1.5

3.0

1.8

10

800

400

1.0

2.5

1.8

HalfSpace

1000

500

1.0

2.0

1.8

Layer
No.

Thickness

Table 6.4

[m]

Medium Properties Used for the Validation of the Inversion Algorithms


(Case 2)

VP

VS

DP

DS

Mass
Density

[m/s]

[m/s]

[%]

[%]

[t/m3]

10

800

400

1.0

3.0

1.8

600

300

1.5

3.5

1.7

10

800

400

1.0

3.0

1.8

HalfSpace

1000

500

1.0

2.5

1.8

Layer
No.

Thickness

[m]

164

Validation of the Algorithms

Table 6.5

Medium Properties Used for the Validation of the Inversion Algorithms


(Case 3)

VP

VS

DP

DS

Mass
Density

[m/s]

[m/s]

[%]

[%]

[t/m3]

1000

500

1.0

2.5

1.8

10

800

400

1.5

3.0

1.8

10

1000

500

1.0

2.5

1.8

HalfSpace

1200

600

1.0

2.0

1.8

Layer
No.

Thickness

[m]

The dispersion curves of the first four modes of propagation for the Case 1 soil profile
are shown in Fig.6.3, together with the synthetic dispersion curve. As expected for this
normally dispersive profile, there is a good agreement between the synthetic dispersion
curve and the dispersion curve associated with the fundamental mode of propagation.

500
Fundamental Mode

Rayleigh Phase Velocity [m/s]

450

Second Mode
Third Mode

400

Fourth Mode
Synthetic Curve

350
300
250
200
150
0

Figure 6.3

20

40

60
80
F requency [Hz]

Rayleigh Dispersion Curves for Case 1 Soil Profile

100

120

Validation of the Algorithms

165

However, Fig. 6.4 also shows that the agreement between the synthetic dispersion curve
and the effective dispersion curve is even better, particularly at frequencies less than 15 Hz. The
effective dispersion curve was computed by averaging the effective Rayleigh phase velocity
V$ R (r , ) over the receiver locations at each frequency. Figure 6.4 shows that, even in
normally dispersive media, the simulated experimental dispersion curve compares better
with the effective dispersion curve than with the modal dispersion curve.

400

Rayleigh Phase Velocity [m/s]

350
Synthetic Curve
Effective Curve
Fundamental Mode

300

250

200

150
0

Figure 6.4

20

40
60
F requency [Hz]

80

100

Rayleigh Effective Dispersion Curve for Case 1 Soil Profile

Figure 6.5 shows modal and synthetic attenuation curves for the Case 1 soil profile. The
irregularities of the synthetic attenuation curve are mostly due to numerical noise affecting
the computation of the attenuation coefficients and caused by the coupling between
material and geometric attenuation.
Because the displacement amplitude decay with distance is largely governed by
geometric attenuation, the evaluation of the Rayleigh attenuation coefficients is sensitive to
numerical noise. Both seismic and electrical noise affect experimental amplitude
measurements. Improvements in the accuracy of amplitude measurements can be achieved
by using noise corrections procedures and other signal processing techniques (Spang, 1995;
Rix et al., 1998a). Figure 6.5 shows that, despite the irregularities, the synthetic attenuation
curve compares fairly well with the attenuation curve associated with the fundamental mode
of propagation. As expected, the modal and synthetic attenuation coefficients increase in
magnitude as the frequency increases.

166

Validation of the Algorithms


0.2

Rayleigh Attenuation Coefficient [1/m]

Fundamental Mode
Second Mode

0.15

Third Mode
Fourth Mode
Synthetic Curve

0.1

0.05

0
0

Figure 6.5

20

40

60
80
Frequency [Hz]

100

120

Rayleigh Attenuation Curves for Case 1 Soil Profile

The modal and synthetic dispersion curves corresponding to the soil profile of Case 2
are illustrated in Fig.6.6, where only the first four modes of propagation have been plotted.

550
Fundamental Mode
Second Mode

Rayleigh Phase Velocity [m/s]

500

Third Mode
Fourth Mode
Synthetic Curve

450

400

350

300
0

Figure 6.6

20

40

60
80
F requency [Hz]

Rayleigh Dispersion Curves for Case 2 Soil Profile

100

120

Validation of the Algorithms

167

Case 2 profile is a soil deposit with a soft layer trapped between two stiffer layers. The
irregularity in soil stratification is reflected in modal dispersion curves which contain, as
shown in Fig.6.6, three jumps discontinuities at frequencies of about 50, 82 and 98 Hz. It is
important to note that modal dispersion curves need not be continuous. Depending on the
medium stratification and on the frequency range, these curves may display one or more
jump discontinuities. The current dispersion curves are computed using a set of 50
frequencies ranging from 5 to 100 Hz. A finer frequency discretization would result in
smoother dispersion curves, but the jump discontinuities are features that are independent
of the frequency discretization. Note that the synthetic dispersion curve also has a jump
discontinuity at a frequency of about 35 Hz.
Figure 6.6 shows that in the frequency range of about 35 to 50 Hz the synthetic
dispersion curve deviates from the fundamental mode dispersion curve; it mostly follows
the second mode of propagation. However, the fundamental mode still dominates the
synthetic dispersion curve for frequencies outside this range.
Figure 6.7 shows a comparison between the effective, the synthetic and the fundamental
mode dispersion curves. The ability of the effective dispersion curve to follow the irregular
pattern of the synthetic dispersion curve over the frequency range of 35-50 Hz is readily
apparent. The results shown in Fig. 6.7 are remarkable. As mentioned earlier in this chapter,
irregular layered media such as the profile of Case 2 are inversely dispersive and their
response is no longer governed by the fundamental mode of propagation as in normally
dispersive media. The synthetic dispersion curve, which simulates the dispersion curve that
would be obtained experimentally, is determined from measurements of displacement
phases. Because the displacement field is obtained from the superposition of all of the
Rayleigh modes of propagation, the synthetic dispersion curve inherently reflects the
consequences of multi-mode wave propagation.
As a result, the agreement between a synthetic or an experimental dispersion curve and
any modal dispersion curve is expected to be poor in media where the response is
controlled by more than one mode of propagation. The effective dispersion curve, on the
other hand, is built on the concept of effective phase velocity, which is the velocity of
propagation of a superposition of harmonic waves with the same frequency and different
wave-number. It is not surprising therefore, that the effective and synthetic dispersion
curves illustrated in Fig.6.7 are almost identical.
The synthetic and the modal attenuation curves associated with the Case 2 soil profile
are shown in Fig.6.8. Because the modal attenuation curves are very similar, it is difficult to
identify a predominant mode of propagation. However, the synthetic attenuation curve
appears to intersect all four modal attenuation curves in several frequency ranges.

168

Validation of the Algorithms


500

Rayleigh Phase Velocity [m/s]

Synthetic Curve

450

Effective Curve
Fundamental Mode

400

350

300
0

Figure 6.7

20

40

60
80
F requency [Hz]

100

120

Rayleigh Effective Dispersion Curve for Case 2 Soil Profile

0.08

Rayleigh Attenuation Coefficient [1/m]

0.07

Fundamental Mode
Second Mode

0.06

Third Mode
Fourth Mode

0.05

Synthetic Curve

0.04
0.03
0.02
0.01
0
0

Figure 6.8

20

40

60
80
Frequency [Hz]

Rayleigh Attenuation Curves for Case 2 Soil Profile

100

120

Validation of the Algorithms

169

The dispersion curves for the Case 3 soil profile are shown in Fig. 6.9. Case 3 is an
example of a soil deposit containing a stiff thin layer that may represent an overconsolidated
crust overlying an otherwise regular soil medium; this soil stratification is encountered at
many geotechnical sites. The modal dispersion curves associated with this soil deposit are
discontinuous, emulating a pattern already observed for the Case 2 soil profile. However, the
modal dispersion curves of Case 3 have only one jump discontinuity that occurs at a
frequency of about 75 Hz.

650
Fundamental Mode
Second Mode

Rayleigh Phase Velocity [m/s]

600

Third Mode
Fourth Mode
Synthetic Curve

550

500

450

400

350
0

Figure 6.9

20

40

60
80
F requency [Hz]

100

120

Rayleigh Dispersion Curves for Case 3 Soil Profile

Two discontinuities, one at about 41 Hz and the other at about 65 Hz, characterize the
synthetic dispersion curve. For frequencies less than 41 Hz, the synthetic curve is in good
agreement with the fundamental mode dispersion curve. However, for frequencies greater
than 41 Hz, the second mode of propagation controls the synthetic dispersion curve.
A comparison between the synthetic, the fundamental mode, and the effective
dispersion curves is shown in Fig. 6.10. As for the Case 1 and Case 2 soil profiles, the
agreement between the synthetic and the effective dispersion curves is excellent. The extent
of the match between the two curves is indicated by the identical values of frequency where
the jump discontinuities occur.
Figure 6.10 confirms the value of the effective dispersion curve as the most appropriate
response function to be compared with the synthetic dispersion curve. Because the

170

Validation of the Algorithms

procedure used to determine the synthetic dispersion curves was the same as that used in
SASW measurements, these findings are expected to be validated by experimental surface
wave data.

650

Rayleigh Phase Velocity [m/s]

600

Synthetic Curve
Effective Curve
Fundamental Mode

550

500

450

400

350
0

Figure 6.10

20

40

60
80
F requency [Hz]

100

120

Rayleigh Effective Dispersion Curve for Case 3 Soil Profile

Rayleigh Attenuation Coefficient [1/m]

0.06

0.05

Fundamental Mode
Second Mode
Third Mode

0.04

Fourth Mode
Synthetic Curve

0.03

0.02

0.01

0
0

Figure 6.11

20

40

60
80
Frequency [Hz]

Rayleigh Attenuation Curves for Case 3 Soil Profile

100

120

Validation of the Algorithms

171

The synthetic and modal attenuation curves for the Case 3 soil profile are illustrated in
Fig.6.11. This figure displays features that are similar to those of the Case 2 soil profile; in
particular note that the modal attenuation curves are closely spaced over the entire
frequency range.
6.3.1

Uncoupled Analyses

6.3.1.1 UFUMA Inversion Algorithms


This section will present the results of the uncoupled inversions of the synthetic
dispersion and attenuation curves for the Case 1, Case 2 and Case 3 soil profiles using the
UFUMA (Uncoupled, Fundamental Mode Analysis) algorithm. Figure 6.12 shows the
sequence of iterations required for the algorithm to converge for the Case 1 soil profile.
From the figure the theoretical dispersion curve corresponding to the first iteration can be
easily identified.

400

Phase Velocity [m/sec]

350
Theoretical
Synthetic

300
Iter. # 7

250

200

150

Iter. # 1

100
0

Figure 6.12

20

40

60
80
F requency [Hz]

100

120

Fundamental Mode Theoretical and Synthetic Dispersion Curves for Case 1


Soil Profile

The theoretical curve associated with the seventh and final iteration is in good
agreement with the synthetic dispersion curve. The corresponding sequence of shear wave
velocity profiles is illustrated in Fig.6.13 where the dashed line indicates the initial model
used in the inversion. The final shear wave velocity profile is shown with a bold line. Finally,
Fig. 6.14 shows the convergence of the algorithm in terms of the Root-Mean-Square (RMS)
error between the synthetic and the theoretical dispersion curve.

172

Validation of the Algorithms


0
5

Depth [m]

10
15
20
Iter. # 7

25
Iter. # 1

30
35
0

Figure 6.13

200

400
600
800
Shear W ave Velocity [m /sec]

1000

1200

Shear Wave Velocity Profile from UFUMA Inversion Algorithm for Case 1
Soil Profile

Exact RMS Misfit

2.5
2
1.5
1
0.5
0
0

Figure 6.14

6
Iteration #

10

12

Convergence of UFUMA Inversion Algorithm for Case 1 Soil Profile

As described in Section 4.3.2, the uncoupled inversion of the experimental attenuation


curve to obtain the shear damping ratio profile D S is a linear problem which can be written

Validation of the Algorithms

173

as GD S = R . In the UFUMA inversion algorithm, G is the matrix formed by the partial


derivatives of the fundamental mode Rayleigh phase velocity with respect to the shear and
compression wave velocities of the soil layers (see Eqs. 3.45 and 3.63b). Figure 6.15 shows
the shear damping ratio profile obtained from the inversion of the equation GD S = R for
the Case 1 soil profile.
0

0.14

0.12

Attenuation Coefficient (1/m)

Theoretical

Depth (m)

10
15
20
25

0.08
0.06
0.04
0.02

30
35

Synthetic

0.1

Figure 6.15

1
2
3
4
Shear Dam ping Ratio (% )

0
0

20

40
60
80
Frequency (Hz)

100

120

Shear Damping Ratio Profile and Theoretical Attenuation Curve from


UFUMA Inversion Algorithm for Case 1 Soil Profile
1

rms Error

0.9
0.8
0.7
0.6
0.5
-10
10

Figure 6.16

10
S m o o t h i n g P a ra m e t e r

10

10

Attenuation Curves RMS Misfit Error using UFUMA Inversion Algorithm


for Case 1 Soil Profile

174

Validation of the Algorithms

Figure 6.15 also compares the synthetic and the theoretical attenuation curves obtained
from the solution of the forward problem GD S = R using the inverted shear damping
ratio profile. The agreement between the two curves is satisfactory and is confirmed by the
RMS error as a function of the Lagrange multiplier shown in Fig.6.16.

500
Theoretical
Synthetic

Phase Velocity [m/sec]

450

400
Iter. # 8

350

300

Iter. # 1

250

200
0

Figure 6.17

20

40

60
80
F requency [Hz]

100

120

Fundamental Mode Theoretical and Synthetic Dispersion Curves for Case 2


Soil Profile

The theoretical and synthetic dispersion curves for the Case 2 soil profile are shown in
Fig.6.17. The non-linear inversion algorithm required eight iterations to converge. However,
the agreement between the synthetic and the theoretical dispersion curve of the final
iteration is not very satisfactory. As mentioned in the previous section, in irregular soil
profiles the effects of higher modes of propagation can no longer be neglected, and
therefore it is unrealistic to expect good agreement between a synthetic (or experimental)
dispersion curve and a theoretical curve containing only the fundamental mode of
propagation. The numerical simulation shown in Fig.6.17 confirms this hypothesis.
The sequence of shear wave velocity profiles during the iterative inversion process is
illustrated in Fig.6.18 where the final profile is again denoted with a bold line. The RMS
error misfit between the synthetic and the simulated dispersion curve as a function of the
iteration number is shown in Fig.6.19.

Validation of the Algorithms

175

0
5
10

Depth [m]

Iter. # 8

15
20
25
Iter. # 1

30
35
0

Figure 6.18

100

200
300
400
500
600
Shear W ave Velocity [m /sec]

700

800

Shear Wave Velocity Profile from UFUMA Inversion Algorithm for Case 2
Soil Profile

4
3.5

Exact RMS Misfit

3
2.5
2
1.5
1
0.5
0
0

Figure 6.19

4
5
Iteration #

Convergence of UFUMA Inversion Algorithm for Case 2 Soil Profile

The results of the uncoupled inversion of the synthetic attenuation curve are shown in
Fig. 6.20 and Fig. 6.21. The agreement between the synthetic and the fundamental mode

176

Validation of the Algorithms

attenuation curves is acceptable but not very satisfactory. This observation is consistent with
the results from the inversion of the synthetic dispersion curve shown in Fig.6.17.
0

0.08

0.07

Attenuation Coefficient (1/m)

Theoretical

Depth (m)

10

15

20

25

30

0.06

Synthetic

0.05
0.04
0.03
0.02
0.01
0

35
3

Figure 6.20

3.25
3.5
3.75
S h e a r D a m p i n g R a tio (% )

20

40
60
80
Frequency (Hz)

100

120

Shear Damping Ratio Profile and Theoretical Attenuation Curve from


UFUMA Inversion Algorithm for Case 2 Soil Profile

rms Error

0.75

0.7

0.65
-10
10

Figure 6.21

10
Smoothing Parameter

10

10

Attenuation Curves RMS Misfit Error using UFUMA Inversion Algorithm


for Case 2 Soil Profile

Validation of the Algorithms

177

The last soil profile to be investigated is Case 3 which is a soil deposit characterized by
the presence of a thin, stiff surface layer overlying an otherwise regular shear wave velocity
profile. Figure 6.22 shows the sequence of theoretical dispersion curves obtained from the
inversion of the synthetic dispersion curve. For this soil profile, the UFUMA inversion
600
Theoretical
Synthetic

Phase Velocity [m/sec]

500
Iter. # 11

400
Iter. # 1

300

200
0

Figure 6.22

20

40

60
80
Frequency [Hz]

100

120

Fundamental Mode Theoretical and Synthetic Dispersion Curves for


Case 3 Soil Profile
0
5

Depth [m]

10
15
20

25
Ite r . # 1
Ite r . # 1 1

30
35
0

Figure 6.23

200
400
600
800
S h e a r W a v e V e lo c i t y [ m / s e c ]

1000

Shear Wave Velocity Profile from UFUMA Inversion Algorithm for Case
3 Soil Profile

178

Validation of the Algorithms

Exact RMS Misfit

0
0

Figure 6.24

6
Iteration #

10

12

Non-Convergence of UFUMA Inversion Algorithm for Case 3 Soil


Profile

algorithm failed to converge. Figure 6.23 shows the shear wave velocity profiles associated
with the theoretical dispersion curves of Fig. 6.22. Figure 6.24 shows the RMS error misfit
as the number of iterations progresses; it is apparent from the oscillatory behavior of the
RMS error that the solution did not converge.
The results obtained for Case 3 soil profile emphasize the limitations of the fundamental
mode approach to the interpretation of surface wave data. In irregular soil deposits where
higher modes of propagation govern the response of the medium, the inversion of the
experimental dispersion curve may be either inaccurate, as occurred in the Case 2 soil
profile, or it may be unable to converge to a solution, as shown for the Case 3 soil profile.
Better results were obtained from the inversion of the synthetic attenuation curve as
illustrated by Fig. 6.25 and Fig. 6.26. A more regular shear damping ratio profile (compared
with the corresponding shear wave velocity profile) is the main reason for this result. In
fact, Fig. 6.11 shows that as a consequence of a smoothly varying D S profile, the modal
attenuation curves for Case 3 are closely spaced, and hence it is not surprising that a
fundamental-mode-based inversion of the equation GD S = R yields satisfactory results.
6.3.1.2 UEQMA Inversion Algorithms
In this section the synthetic dispersion and attenuation curves for the Case 1, Case 2 and
Case 3 soil profiles are inverted to obtain the shear wave velocity and shear damping ratio
profiles using the UEQMA (Uncoupled, Equivalent-Multi-Mode Analysis) algorithm. As
described in Section 4.3.3, this inversion algorithm accounts for all the Rayleigh modes of
propagation via the concept of effective phase velocity.

Validation of the Algorithms

179

Fig.6.27 shows the synthetic and the theoretical dispersion curves for the Case 1 soil
profile. Only five iterations were required for the algorithm to converge. From Fig.6.27 it
can be seen that the agreement between the theoretical effective and the synthetic
dispersion curves is excellent for the final iteration.
0.05

0.045

5
Theoretical

0.04

Synthetic

Attenuation Coefficient (1/m)

Depth (m)

10

15

20

25

0.035
0.03
0.025
0.02
0.015
0.01

30
0.005

35
2

Figure 6.25

2.5
3
3.5
Shear Damping Ratio (% )

0
0

20

40

60
80
Frequency (Hz)

100

120

Shear Damping Ratio Profile and Theoretical Attenuation Curve from


UFUMA Inversion Algorithm for Case 3 Soil Profile
0.4

rms Error

0.35

0.3

0.25
-10
10

Figure 6.26

10

-5

10
10
Smoothing Parameter

10

10

Attenuation Curves RMS Misfit Error using UFUMA Inversion Algorithm


for Case 3 Soil Profile

180

Validation of the Algorithms


400

Phase Velocity [m/sec]

350
Theoretical
Synthetic
300

250
Iter. # 5

200
Iter. # 1

150

100
0

Figure 6.27

20

40

60
80
Frequency [Hz]

100

120

Effective Theoretical and Synthetic Dispersion Curves for Case 1 Soil


Profile

The sequence of shear wave velocity profiles is shown in Fig.6.28 where the dashed line
is used to denote the starting profile, and the bold line corresponds to the final profile.
0
5

Depth [m]

10
15
20
25
Iter. # 5

30

Iter. # 1

35
0

Figure 6.28

200

400
600
800
Shear W ave Velocity [m /sec]

1000

1200

Shear Wave Velocity Profile from UEQMA Inversion Algorithm for Case 1
Soil Profile

Validation of the Algorithms

181

The RMS error misfit between the theoretical effective and the synthetic dispersion
curves as a function of the iteration number is shown in Fig.6.29. In the UEQMA inversion
algorithm the shear damping ratio profile is obtained from the inversion of experimental
$ is formed by the partial
$ S = R . The matrix G
attenuation curve via the equation GD
derivatives of the effective Rayleigh phase velocity with respect to the shear and
compression wave velocities of the soil layers averaged over the receiver spacings (see Eq.
3.58).
3

Exact RMS Misfit

2.5
2
1.5
1
0.5
0

Figure 6.29

6
Iteration #

10

12

Convergence of UEQMA Inversion Algorithm for Case 1 Soil Profile

0.14

0.12
Theoretical

Attenuation Coefficient (1/m)

Depth (m)

10

15

20

25

30

35
2.5

Figure 6.30

Synthetic

0.1

0.08

0.06

0.04

0.02

3
3.5
4
4.5
Shear Damping Ratio (%)

0
0

20

40

60
80
Frequency (Hz)

100

120

Shear Damping Ratio Profile and Theoretical Attenuation Curve from


UEQMA Inversion Algorithm for Case 1 Soil Profile

182

Validation of the Algorithms


1.1

rms Error

0.9

0.8

0.7
-10
10

Figure 6.31

10
S m oothing Param eter

10

10

Attenuation Curves RMS Misfit Error using UEQMA Inversion Algorithm


for Case 1 Soil Profile

Figure 6.30 shows the resulting shear damping ratio profile and theoretical attenuation
curve for the Case 1 soil profile. At high frequencies the theoretical attenuation curve
becomes very irregular emulating the synthetic attenuation curve (see also Fig.6.5). The RMS
error misfit of the synthetic and predicted attenuation curve is illustrated in Fig.6.31.
The results for the Case 2 soil profile are presented in Fig.6.32, which shows the
succession of theoretical effective dispersion curves required for the UEQMA inversion
algorithm to converge. The results obtained are remarkable: the cusped synthetic dispersion
curve is almost perfectly matched by the theoretical curve corresponding to the final
iteration. Figure 6.32 also shows the stability and convergence of the inversion algorithm.
As illustrated in the previous section, the fundamental-mode-based approach was unable
to yield a theoretical dispersion curve that matches the cusped synthetic curve well.
Conversely, this difficult task has been successfully solved by the effective, multi-mode
approach of the UEQMA algorithm. The sequence of shear wave velocity profiles
associated with the theoretical dispersion curves of Fig. 6.32 is illustrated in Fig. 6.33. The
final profile is denoted using a bold line.
The convergence of the algorithm in terms of the RMS error misfit between theoretical
and synthetic dispersion curves is shown in Fig. 6.34. Ten iterations were required for the
algorithm to converge.

Validation of the Algorithms

183

Phase Velocity [m/sec]

600

500

Iter. # 10

400

Iter. # 1

300

200

Figure 6.32

Theoretical
S ynthetic

20

40
60
80
F requency [Hz]

100

120

Effective Theoretical and Synthetic Dispersion Curves for Case 2 Soil


Profile

Depth [m]

10

15

Iter. # 10

20

25
Iter. # 1

30

35
0

Figure 6.33

200
400
600
S h e a r W a v e V e lo c i t y [ m / s e c ]

800

Shear Wave Velocity Profile from UEQMA Inversion Algorithm for Case 2
Soil Profile

184

Validation of the Algorithms


3

Exact RMS Misfit

2.5
2
1.5
1
0.5
0
0

Figure 6.34

6
Iteration #

10

12

Convergence of UEQMA Inversion Algorithm for Case 2 Soil Profile

Figures 6.35 and 6.36 summarize the results obtained from the inversion of the synthetic
attenuation curve. The agreement between the theoretical and the synthetic attenuation
curves is satisfactory.
0

0.06

0.05

Synthetic

Attenuation Coefficient (1/m)

10

Depth (m)

Theoretical

15

20

25

0.04

0.03

0.02

0.01

30

35
3

Figure 6.35

3.2

3.4
3.6
3.8
Shear Damping Ratio (%)

0
0

20

40

60
80
Frequency (Hz)

100

120

Shear Damping Ratio Profile and Theoretical Attenuation Curve from


UEQMA Inversion Algorithm for Case 2 Soil Profile

Validation of the Algorithms

185

0.66

rms Error

0.65
0.64
0.63
0.62
0.61
-10
10

Figure 6.36

10
Smoothing Parameter

10

10

Attenuation Curves RMS Misfit Error using UEQMA Inversion Algorithm


for Case 2 Soil Profile

Figure 6.37 illustrates the sequence of theoretical dispersion curves successfully


converging to the synthetic dispersion curve for the Case 3 soil profile. Figure 6.37 shows
that the performance of the inversion algorithm UEQMA for the Case 3 soil profile is also
remarkable. It should be recalled that the UFUMA algorithm failed to converge for Case 3.
With the UEQMA algorithm, convergence is achieved after only six iterations.
700
Theoretical
Synthetic

Phase Velocity [m/sec]

600

500
Iter. # 6

400

300

Iter. # 1

200
0

Figure 6.37

20

40

60
80
Frequency [Hz]

100

120

Effective Theoretical and Synthetic Dispersion Curves for Case 3 Soil


Profile

186

Validation of the Algorithms

Figure 6.38 illustrates the sequence of shear wave velocity profiles corresponding to the
six iterations. The RMS error misfit between the theoretical and the synthetic dispersion
curves as a function of the iteration number is shown in Fig. 6.39. Finally, the results of the
linear inversion of the synthetic attenuation curve are presented in Fig.6.40 and Fig.6.41.

0
5

Iter. # 6

Depth [m]

10
15
20
25
Iter. # 1

30
35
0

Figure 6.38

200

400
600
Shear W ave Velocity [m /sec]

800

1000

Shear Wave Velocity Profile from UEQMA Inversion Algorithm for Case 3
Soil Profile

Exact RMS Misfit

0
0

Figure 6.39

3
4
Iteration #

Convergence of UEQMA Inversion Algorithm for Case 3 Soil Profile

Validation of the Algorithms

187
0.05

0.045
5

Theoretical

0.04

Attenuation Coefficient (1/m)

Depth (m)

10

15

20

25

Synthetic

0.035
0.03
0.025
0.02
0.015
0.01

30
0.005
35
2

Figure 6.40

2.5
3
Shear Damping Ratio (%)

3.5

0
0

20

40

60
80
Frequency (Hz)

100

120

Shear Damping Ratio Profile and Theoretical Attenuation Curve from


UEQMA Inversion Algorithm for Case 3 Soil Profile

0.6

rms Error

0.55

0.5

0.45

0.4

0.35
-10
10

Figure 6.41

10
S m o o t h in g P a ra m e t e r

10

10

Attenuation Curves RMS Misfit Error using UEQMA Inversion Algorithm


for Case 3 Soil Profile

188
6.3.2

Validation of the Algorithms


Coupled Analyses

6.3.2.1 CFUMA Inversion Algorithms


In the previous two sections, the inversion of the synthetic dispersion and attenuation
curves was performed independently using the fundamental mode (Section 6.3.1.1) or the
multi-mode approach (Section 6.3.1.2). However, in this and the next sections, the
dispersion and attenuation curves are inverted simultaneously. In particular, this section will
illustrate the results of the coupled inversion of these curves for Case 1, Case 2 and Case 3
soil profiles using the CFUMA (Coupled, Fundamental Mode Analysis) algorithm.
Figure 6.42 shows the sequence of fundamental mode dispersion and attenuation curves
for Case 1 soil profile. The agreement between the theoretical dispersion curve
corresponding to the sixth and final iteration and the synthetic dispersion curve is good for
frequencies greater than 20 Hz. However, at frequencies lower than 20 Hz the inversion
algorithm was unable to obtain a satisfactory match between these two curves.

Attenuation Coefficient [1/m]

Phase Velocity [m/sec]

600

Figure 6.42

Theoretical

500
400

Synthetic

Iter. # 1

300
Iter. # 6

200
100
0

20

40
60
80
Frequency [Hz]

100

120

0.2
Theoretical

0.15

Iter. # 6

Synthetic

0.1
0.05
0
0

Iter. # 1

20

40
60
80
Frequency [Hz]

100

120

Fundamental Mode Theoretical Dispersion and Attenuation Curves for Case


1 Soil Profile

Validation of the Algorithms

189

The agreement between the synthetic and theoretical attenuation curves for the final
iteration is good at all frequencies. Figure 6.43 shows the sequence of shear wave velocity
and shear damping ratio profiles corresponding to the dispersion and attenuation curves of
Fig. 6.42. The dashed and the bold lines denote the initial and final profiles, respectively.

0
Iter. # 1

Iter. # 1

10

15

15

Depth [m]

Depth [m]

Iter. # 6

10

20
25

25

30

30

Iter. # 6

35

0
200 400 600 800
Shear Wave Velocity [m/sec]

Figure 6.43

20

35

0
2
4
6
Shear Damping Ratio [%]

Shear Wave Velocity and Shear Damping Ratio Profile from CFUMA
Inversion Algorithm for Case 1 Soil Profile
12

Exact RMS Misfit

10
8
6
4
2
0
0

Figure 6.44

6
Iteration #

10

12

Convergence of CFUMA Inversion Algorithm for Case 1 Soil Profile

190

Validation of the Algorithms

A measure of the goodness of the simultaneous inversion is provided also in this case by
the RMS error misfit between the complex valued theoretical and synthetic dispersion
curves. Figure 6.44 shows the RMS error as a function of the iteration number for Case 1
soil profile. The inversion algorithm required six iterations to converge.
The results of the simultaneous inversion of the synthetic dispersion and attenuation
curves for Case 2 are presented in Figs. 6.45 through 6.47. In particular, Fig. 6.45 shows the
theoretical and synthetic dispersion and attenuation curves for this soil profile. The
agreement between the synthetic dispersion and attenuation curves and the corresponding
theoretical curves of the sixth and last iteration is satisfactory. By comparing Fig. 6.45 with
Figs.6.17 and 6.20, it appears that for the fundamental mode-based algorithms, the
simultaneous inversion furnishes more accurate results than the uncoupled inversion for this
soil profile. It should also be noted that the simultaneous inversion required a fewer number
of iterations to converge than the corresponding uncoupled inversion (six vs. eight).

Attenuation Coefficient [1/m]

Phase Velocity [m/sec]

500

Figure 6.45

Iter. # 6

400
300

Iter. # 1

Theoretical
Synthetic

200
100

20

40
60
80
Frequency [Hz]

100

120

0.1

Theoretical
Synthetic

0.05

Iter. # 1
Iter. # 6

20

40
60
80
Frequency [Hz]

100

120

Fundamental Mode Theoretical Dispersion and Attenuation Curves for Case


2 Soil Profile

Validation of the Algorithms

191

Figure 6.46 shows the sequence of shear wave velocity and shear damping ratio obtained
from the simultaneous inversion as the number of iteration progresses, and finally Fig. 6.47
illustrates the convergence of the algorithm in terms of the RMS error misfit.

Iter. # 6

10

10

15

15

Depth [m]

Depth [m]

Iter. # 6

20
25

20
25
Iter. # 1

Iter. # 1

30

30

35

35

0
200 400 600 800
Shear Wave Velocity [m/sec]

Figure 6.46

0
2
4
6
Shear Damping Ratio [%]

Shear Wave Velocity and Shear Damping Ratio Profile from CFUMA
Inversion Algorithm for Case 2 Soil Profile

Exact RMS Misfit

0
0

Figure 6.47

4
6
Iteration #

10

Convergence of CFUMA Inversion Algorithm for Case 2 Soil Profile

192

Validation of the Algorithms

Theoretical and synthetic dispersion and attenuation curves for Case 3 are shown in Fig.
6.48. The agreement between the synthetic and theoretical curves for the last iteration is
satisfactory. Recall that the fundamental mode, uncoupled inversion of the synthetic
dispersion curve had failed to converge.

Attenuation Coefficient [1/m]

Phase Velocity [m/sec]

600

Figure 6.48

Theoretical
Synthetic

500

Iter. # 8

400
Iter. # 1

300

20

40

60
80
Frequency [Hz]

100

120

0.1
Theoretical
Synthetic

0.05
Iter. # 1
Iter. # 8

0
0

20

40

60
80
Frequency [Hz]

100

120

Fundamental Mode Theoretical Dispersion and Attenuation Curves for Case


3 Soil Profile

The sequence of shear wave velocity and shear damping ratio profiles for Case 3 is
illustrated in Fig. 6.49. Finally, the RMS error misfit between the synthetic dispersion and
attenuation curves is shown in Fig.6.50. The inversion algorithm required eight iterations to
converge.
6.3.2.2 CEQMA Inversion Algorithms
The CEQMA inversion algorithm combines the features of the simultaneous inversion
with those of the multi-mode propagation via the concept of effective velocity. This is the
most sophisticated of the inversion algorithms developed in this study. Unfortunately, the
results obtained thus far using this technique indicate only moderate success. It is believed
that most of the problems exhibited by the algorithm are due to an inability of the

Validation of the Algorithms

193

algorithm to determine, at certain frequencies, the correct sequence of Rayleigh modes


(particularly the higher modes).
0

0
Iter. # 8

10

10

15

15

Iter. # 8

20
25

Depth [m]

Depth [m]

Iter. # 1

20
25

Iter. # 1

30

30

35
35
200
400
600
800
0
2
4
6
Shear Wave Velocity [m/sec] Shear Damping Ratio [%]

Figure 6.49

Shear Wave Velocity and Shear Damping Ratio Profile from CFUMA
Inversion Algorithm for Case 3 Soil Profile

Exact RMS Misfit

Figure 6.50

6
Iteration #

10

12

Convergence of CFUMA Inversion Algorithm for Case 3 Soil Profile

194

Validation of the Algorithms

Because the strategy of the CEQMA inversion algorithm is based on matching the
complex-valued effective phase velocity with the complex-valued synthetic dispersion curve,
an error in computing the correct sequence of modes (even at one frequency) will ultimately
cause the effective phase velocity to be incorrect. Figures 6.51 through 6.53 shows the
results for Case 1 obtained from the application of the CEQMA algorithm.

Attenuation Coefficient [1/m]

Phase Velocity [m/sec]

600

Figure 6.51

Theoretical

500

Synthetic

400

Iter. # 1

300
Iter. # 6

200
100
0

20

40
60
80
Frequency [Hz]

100

120

0.2
Theoretical

0.15

Synthetic

Iter. # 6

0.1
0.05
0
0

Iter. # 1

20

40
60
80
Frequency [Hz]

100

120

Effective Theoretical Dispersion and Attenuation Curves for Case 1 Soil


Profile

The instability of the algorithm is indicated by irregular pattern of the effective


dispersion and attenuation curves at several iterations. The sequence of shear wave velocity
and shear damping ratio profiles is illustrated in Fig. 6.52. Finally, Fig. 6.53 shows the RMS
error misfit.
The results of the simultaneous inversion of surface wave data for Case 2 soil profile are
shown in Figs.6.54 through 6.56. For this soil profile, the theoretical dispersion and
attenuation curves also exhibit an irregular pattern, particularly in the case of the attenuation
curves.

Validation of the Algorithms

195

0
Iter. # 1

10

10

15

15

20

Depth [m]

Depth [m]

Iter. # 1

Iter. # 6

Iter. # 6

20

25

25

30

30

35

35
0
200
400
600
0
2
4
6
Shear Wave Velocity [m/sec] Shear Damping Ratio [%]

Figure 6.52

Shear Wave Velocity and Shear Damping Ratio Profile from CEQMA
Inversion Algorithm for Case 1 Soil Profile

12

Exact RMS Misfit

10
8
6
4
2
0
0

Figure 6.53

6
Iteration #

10

12

RMS Error Misfit of CEQMA Inversion Algorithm for Case 1 Soil Profile

196

Validation of the Algorithms

Attenuation Coefficient [1/m]

Phase Velocity [m/sec]

600

Figure 6.54

Theoretical

500

Synthetic

400

Iter. # 7

300
Iter. # 1

200
0

20

40

60
80
Frequency [Hz]

120

0.1
Theoretical

Iter. # 7

Synthetic

0.05
Iter. # 1

0
0

20

40

60
80
Frequency [Hz]

100

120

Effective Theoretical Dispersion and Attenuation Curves for Case 2 Soil


Profile
0

0
Iter. # 7

Iter. # 1

10

15

15

Depth [m]

Depth [m]

10

20

Iter. # 7

Iter. # 1

Figure 6.55

100

20

25

25

30

30

35
200
400
600
Shear Wave Velocity [m/sec]

35
0
2
4
6
8
Shear Damping Ratio [%]

Shear Wave Velocity and Shear Damping Ratio Profile from CEQMA
Inversion Algorithm for Case 2 Soil Profile

Validation of the Algorithms

197

Exact RMS Misfit

0
0

Figure 6.56

6
Iteration #

10

12

RMS Error Misfit of CEQMA Inversion Algorithm for Case 2 Soil Profile

The final results are those for the Case 3 soil profile that are presented in Figs. 6.57
through 6.59. They are characterized by the same irregular behavior of the previous two soil
profiles, which is also reflected by the oscillatory behavior of the RMS error misfit.

Attenuation Coefficient [1/m]

Phase Velocity [m/sec]

600

Figure 6.57

Iter. # 6

500
400

Iter. # 1

300

Theoretical

200

Synthetic

100

20

40

60
80
Frequency [Hz]

100

120

0.1
Theoretical
Synthetic
0.05
Iter. # 1
Iter. # 6

20

40

60
80
Frequency [Hz]

100

120

Effective Theoretical Dispersion and Attenuation Curves for Case 3 Soil


Profile

198

Validation of the Algorithms


0

5
Iter. # 6

10

15

Depth [m]

Depth [m]

10

20

15
Iter. # 6

20

Iter. # 1

25

25

30

30
Iter. # 1

35

0
200 400 600 800
Shear W ave Velocity [m/sec]

Figure 6.58

35

0
2
4
6
Shear Damping Ratio [%]

Shear Wave Velocity and Shear Damping Ratio Profile from CEQMA
Inversion Algorithm for Case 3 Soil Profile

Exact RMS Misfit

0
0

Figure 6.59

6
Iteration #

10

12

RMS Error Misfit of CEQMA Inversion Algorithm for Case 3 Soil Profile

Validation of the Algorithms


6.3.3

199

Results and Discussion

This section will summarize and discuss the results obtained from the inversion of the
synthetic surface wave data for the Case 1, Case 2, and Case 3 soil profiles. One should not
expect, even in a numerical simulation, to obtain shear wave velocity and shear damping
ratio profiles matching the original stratigraphy exactly. This is because the numerical
simulation was performed using the same data reduction procedures employed in
experimental SASW measurements. Therefore, the inversion of synthetic dispersion and
attenuation curves will inevitably suffer the limitations that are inherent in the SASW
methodology, particularly the problem of non-uniqueness (see Chapter 4).
Nevertheless, the four algorithms UFUMA, UEQMA, CFUMA, and CEQMA are based
on different inversion strategies and it is instructive to compare and evaluate their
performance. The results of the algorithm CEQMA have not been compared with those of
the other algorithms because, based on the discussion in the previous section, they are
considered unreliable. Figures 6.60 and 6.61 show the shear wave velocity and shear
damping ratio profiles corresponding to Case 1.
The RMS error misfit between the original and the predicted VS and D S profiles, which is
defined by VSorig VSpred

D pred
n L and D orig
S
S

n L respectively (nL is the number

of layers), was also computed. The results of this calculation for the four algorithms are
reported in Table 6.6.
0
UFUMA

UEQMA
CFUMA
O R IG I N A L

Depth [m]

10
15
20
25
30
35
100

Figure 6.60

200

300
400
500
Shear W ave Velocity [m /sec]

600

Inverted Shear Wave Velocity Profiles for Case 1 Soil Stratigraphy

200

Validation of the Algorithms


0
UFUMA

UEQMA
CFUMA
O R IG I N A L

Depth [m]

10
15
20
25
30
35

2
3
4
Shear Damping Ratio [% ]

Figure 6.61

Inverted Shear Damping Ratio Profiles for Case 1 Soil Stratigraphy

Table 6.6

Inversion Algorithms RMS Error Misfit for Case 1 Soil Profile

UFUMA

UEQMA

CFUMA

VS

57.1

37.9

46.8

DS

0.66

1.01

0.45

From this table it appears that the algorithms providing the most accurate predictions of
the original profile are the uncoupled, multi-mode analysis (UEQMA) and coupled,
fundamental mode inversion (CFUMA). In terms of the shear wave velocity profile, the
UEQMA algorithm is the most accurate (i.e., the lowest RMS error misfit); the uncoupled
fundamental mode-based UFUMA algorithm yielded the worst prediction. In terms of the
shear damping ratio profile the lowest RMS error was attained by the coupled inversion
with the CFUMA algorithm.

Validation of the Algorithms

201

For the Case 2 soil profile, the results of the inversions are summarized by Figs. 6.62 and
6.63 and by Table 6.7. Case 2 is characterized by an irregular stratigraphy where a soft layer is
trapped between two harder layers. As mentioned several times in this chapter, such layered
media are inversely dispersive and their response includes contributions from higher modes
of propagation. This feature is reflected in the performance of the inversion algorithms.
0
UFUMA

UEQMA
CFUMA

Depth [m]

10

ORIGINAL

15
20
25
30
35
200

Figure 6.62

300
400
500
S h e a r W a v e V e l o c i t y [ m /sec]

600

Inverted Shear Wave Velocity Profiles for Case 2 Soil Stratigraphy


0
5

Depth [m]

10
15
20

UFUMA
UEQMA
CFUMA

25

ORIGINAL

30
35
0

Figure 6.63

4
6
Shear Damping Ratio [%]

10

Inverted Shear Damping Ratio Profiles for Case 2 Soil Stratigraphy

202
Table 6.7

Validation of the Algorithms


Inversion Algorithms RMS Error Misfit for Case 2 Soil Profile

UFUMA

UEQMA

CFUMA

VS

18.5

14.0

38.3

DS

0.72

0.62

1.72

The lowest RMS error misfit for both shear wave velocity and shear damping ratio
profiles was achieved by the effective, multi-mode analysis of the UEQMA algorithm. The
results obtained by the fundamental mode based algorithms UFUMA and CFUMA were in
both cases less accurate, particularly those associated with the algorithm CFUMA. This
result confirms the inadequacy of the fundamental mode approach for irregular media, and
reinforces the need for an approach based on multi-mode wave propagation such as the
effective Rayleigh phase velocity.
Finally, Figs. 6.64 and 6.65 and Table 6.8 illustrate the results for the Case 3 soil profile.
For this irregular stratigraphy characterized by the presence of a thin and stiff surface crust,
the analyses based on the effective Rayleigh phase velocity again provided the best overall
results. The multi-mode UEQMA inversion algorithm exhibited the best performance in
terms of RMS error misfit for the prediction of the shear wave velocity profile. The
performance of the uncoupled fundamental mode algorithm UFUMA was the least
accurate. In terms of the prediction of shear damping ratio profile, both the UEQMA and
UFUMA algorithms performed well, whereas CFUMA algorithm exhibited a higher RMS
error.
Table 6.9 summarizes the performance of the algorithms UFUMA, CFUMA, and
UEQMA in this numerical simulation. Even though more numerical simulations are
required for a definitive validation of these algorithms, the results obtained in this section
are important, and they lead to the following observations. Overall, the multi-mode-based
approach of the algorithm UEQMA is resulted to be the most accurate inversion algorithm,
particularly for determining the shear wave velocity profile. The fundamental mode-based
algorithms UFUMA and CFUMA are generally less accurate than the algorithm UEQMA in
predicting both the shear wave velocity and the shear damping ratio profiles. There are
however two exceptions to this general rule. The first exception is the prediction of the
shear damping ratio for Case 1 soil profile. Here the algorithm CFUMA has yielded the most
accurate results, whereas the algorithm UEQMA yielded the highest RMS error misfit. The

Validation of the Algorithms

203

second exception is the prediction of shear damping ratio for Case 3 soil profile where both
UFUMA and UEQMA algorithms yielded approximately the same RMS error.
0
UFUMA

UEQMA
CFUMA

Depth [m]

10

ORIGINAL

15
20
25
30
35
300

Figure 6.64

400
500
600
Shear W ave Velocity [m/sec]

700

Inverted Shear Wave Velocity Profiles for Case 3 Soil Stratigraphy

0
UFUMA

UEQMA
CFUMA

Depth [m]

10

ORIGINAL

15
20
25
30
35

Figure 6.65

2
3
4
5
Shear Damping Ratio [%]

Inverted Shear Damping Ratio Profiles for Case 3 Soil Stratigraphy

204
Table 6.8

Table 6.9

Validation of the Algorithms


Inversion Algorithms RMS Error Misfit for Case 3 Soil Profile

UFUMA

UEQMA

CFUMA

VS

47.3

5.6

18.9

DS

0.51

0.54

1.03

Inversion Algorithms Performance in Terms of RMS Error Misfit

VS

DS

Min [RMS]

Max [RMS]

Min [RMS]

Max [RMS]

Case 1

UEQMA

UFUMA

CFUMA

UEQMA

Case 2

UEQMA

CFUMA

UEQMA

CFUMA

Case 3

UEQMA

UFUMA

UFUMA/
UEQMA

CFUMA

It should be remarked that both exceptions occurred in the prediction of the sheardamping ratio D S . This is not surprising since a) the attenuation properties of a dissipative
medium affect the overall response of a medium to a lesser extent than stiffness, and b) in the
boundary value problem considered in this study (i.e. a harmonic point load over the surface
of a vertically heterogeneous viscoelastic half-space) material attenuation is coupled with
geometric attenuation. The combination of the effects of a) and b) makes the computation
of the Rayleigh attenuation coefficients and therefore, the shear damping ratio, very
sensitive to numerical noise.

Validation of the Algorithms

205

Although the numerical simulation presented in this study is not statistically significant
and should be substantiated with other simulations using different profiles and material
parameters, it is still possible to draw two general conclusions. The first and most important
is that a multi-mode inversion analysis yields more accurate results than a fundamental mode
analysis. This is particularly true for irregular media (i.e. Case 2 and Case 3), and for
predictions of the shear wave velocity structure.
The second conclusion is that in regular soil deposits (i.e. Case 1) a fundamental mode
analysis may yield sufficiently accurate results for both shear wave velocity and shear
damping ratio determination. However, the results obtained from the numerical simulation
seem also to indicate that the coupled inversion (more precisely, the strongly coupled
inversion), where the shear wave velocity and shear damping ratio are determined
simultaneously (CFUMA), is more accurate than the corresponding uncoupled analysis
(UFUMA).

206

Validation of the Algorithms

7 EXPERIMENTAL RESULTS

7.1 Overview
The algorithms developed in this research study for the interpretation of surface wave
measurements will now be applied to experimental data. Rayleigh phase velocity and
attenuation measurements were performed at the Treasure Island National Geotechnical
Experimentation Site (California, USA), where independent in situ and laboratory
measurements of shear wave velocity and shear damping ratio are available for comparison.
After a brief description of the geotechnical characteristics of the site, the following
sections will illustrate the results of both the coupled and uncoupled inversion of the
experimental dispersion and attenuation curves. The chapter ends with a comparison of
these results with those obtained from independent cross-hole and laboratory
measurements performed at the site.
7.2 Treasure Island Naval Station Site
Treasure Island is a man-made island constructed of hydraulic fill soils in the eastern
portion of San Francisco Bay. The area investigated for this study is along the western
property margin of the National Geotechnical Experimentation Site (NGES) (see Figure
7.1). Extensive geotechnical characterization of the site was performed for the EPRI (1993)
N
Parking
Fire Station
Building 157

Storage

B1
B2
B3

B4 B5

Current Test Area

EPRI Soil Borings


Note: Not to scale

Figure 7.1

Treasure Island National Geotechnical Experimentation Site (After Spang, 1995)


207

208

Experimental Results

study and included soil borings, penetration tests, and in situ seismic tests. The approximate
locations of the borings used for cross-hole tests in the EPRI study are shown in Figure 7.1.
The soil conditions consist of approximately 10 m of loose, fine to medium sand
underlain by 12 to 18 m of soft clay (Bay Mud). Beneath the Bay Mud are stiffer soils (Older
Bay Mud) which are underlain at depths on the order of 80 m by bedrock. The water table
is approximately 1.5 m below the ground surface. A generalized soil profile of the Treasure
Island site is presented in Figure 7.2. The cone penetration (CPT) and standard penetration
(SPT) test data shown in Figure 7.2 were obtained from the EPRI study.
Cone Tip Resistance (MPa) SPT N Value
0
10
20
0 5 10 15

Soil Profile

0
Gravelly
Sand

Depth (m)

Fine

6
To
Medium

Sand

(Bay Mud)

12

Figure 7.2

Soft Clay

Soil Profile and Properties at the Treasure Island NGES (After Spang, 1995)

Surface wave tests were performed at Treasure Island NGES to determine the
experimental dispersion and attenuation curves at the site. Rayleigh waves were generated by
a vertically oscillating, electrodynamic force generator operating in swept-sine mode. The
dynamic force provided by the shaker is frequency-dependent, with the maximum dynamic

Experimental Results

209

force provided at lower frequencies. Frequencies used in field tests ranged from 5 to 100
Hz, so that the dynamic force supplied by the electromechanical ranged from about 90 to
500 N. Rayleigh waves were recorded by vertical transducers (geophones) having a natural
frequency of 1 Hz at various offsets (r) from the source.
The geophones were spaced at 1.5-m intervals up to an offset of 9 m, at 3-m intervals
from 9 to 30 m, and at 6-m intervals from 30 m to the maximum offset of 60 m. Ten
particle velocity spectra were averaged in the frequency domain at each geophone offset.
The acceleration of the shaker mass was measured with a piezoelectric accelerometer to
ensure that the source was repeatable throughout the test. Particle velocity spectra were
corrected in order to mitigate the effects of ambient noise on attenuation measurements.
The experimental dispersion and attenuation curves were obtained using the
conventional techniques described in Section 5.2. In particular, the Rayleigh phase velocity
was determined using the so-called two-station method, whereas the Rayleigh attenuation
coefficients were calculated with the multi-station method and using a geometric spreading
function accounting for multi-mode Rayleigh wave propagation.
7.3 Uncoupled Inversion
This section will present the results of the uncoupled inversion of the experimental
dispersion and attenuation curves obtained at Treasure Island NGES. The stratigraphy used
for the inversion analyses was selected based on the interpretation of the geotechnical data
shown in Fig.7.2.
In this section, the results associated with the fundamental mode analysis will be
illustrated first (i.e. the UFUMA algorithm), followed by those based on the equivalent
multi-mode analysis (i.e. the UEQMA algorithm). Figure 7.3 shows the sequence of
dispersion curves required for the UFUMA algorithm to converge.
A homogeneous profile with VS = 60 m / s was used as the initial estimate of the layer
shear wave velocities. Five iterations were required for the algorithm to converge. Figure 7.4
shows the corresponding sequence of shear wave velocity profiles obtained from the
fundamental-mode-based inversion. The dashed line indicates the initial profile whereas the
bold line shows the final shear wave velocity profile. The Root-Mean-Square (RMS) error
between the experimental and the theoretical dispersion curve as function of the iteration
number is illustrated in Fig.7.5.

210

Experimental Results
250
Theoretical
Experimental

Phase Velocity [m/sec]

200

150
Iter. # 5

100

50

Iter. # 1

0
0

Figure 7.3

10

20

30
40
50
Frequency [Hz]

60

70

80

Fundamental Mode Theoretical and Experimental Dispersion Curves at


Treasure Island NGES
0
Iter. # 5

Depth [m]

10
Iter. # 1

15
0

Figure 7.4

50

100
150
200
Shear Wave Velocity [m/sec]

250

Shear Wave Velocity Profile from UFUMA Inversion Algorithm at Treasure


Island NGES

Experimental Results

211

8
7

Exact RMS Misfit

6
5
4
3
2
1
0
0

10

Iteration #

Figure 7.5

Convergence of UFUMA Inversion Algorithm at Treasure Island NGES

The results of the fundamental mode uncoupled inversion of the experimental


attenuation curve are presented in Figs. 7.6 and Fig.7.7. In particular, Fig. 7.6 illustrates the
shear damping ratio profile, and the theoretical attenuation curve obtained from the
solution of the forward problem GD S = R .
The RMS error between the theoretical and the experimental attenuation curves as a
function of the smoothing parameter is shown in Fig.7.7. The value selected for the
smoothing parameter is the smallest value that yields a solution vector composed of nonnegative shear damping ratios. For this site the adopted value was 2.8 10 4 which resulted in
a RMS error of 1.38.
The results of the uncoupled equivalent multi-mode analysis conducted with the
algorithm UEQMA will now be illustrated. Figure 7.8 shows the experimental dispersion
curve and the succession of theoretical dispersion curves as the iterations progress. The
same homogeneous profile with VS = 60 m / s was assumed as the initial estimate of the
shear wave velocities.
Figure 7.9 shows the sequence of shear wave velocity profiles obtained from the nonlinear inversion. The bold line indicates the final shear wave velocity profile. The algorithm
required five iterations to converge. The RMS error misfit between the experimental and
the theoretical dispersion curves as a function of the iteration number is shown in Fig.7.10.

212

Experimental Results
0.05

Experimental
Theoretical

Attenuation Coefficient [1/m]

0.04

Depth [m]

10

0.03

0.02

0.01

0.00

15
0.0

Figure 7.6

1.0
2.0
Shear Damping Ratio [%]

3.0

20

40

60

80

Frequency [Hz]

Shear Damping Ratio Profile and Theoretical Attenuation Curve from


UFUMA Inversion Algorithm at Treasure Island NGES

4.0
3.5

Rms Error

3.0
2.5
2.0
1.5

1.38

1.0
0.5
0.0
1.E-06

1.E-03 1.E+00 1.E+03 1.E+06 1.E+09 1.E+12


Smoothing Parameter

Figure 7.7

Attenuation Curves RMS Error using UFUMA Inversion Algorithm at


Treasure Island NGES

Experimental Results

213

250
Theoretical
Experimental

Phase Velocity [m/sec]

200

150
Iter. # 5

100

50

Figure 7.8

Iter. # 1

10

20

30
40
50
Frequency [Hz]

60

70

80

Effective Theoretical and Experimental Dispersion Curves at Treasure


Island NGES

0
Iter. # 5

Depth [m]

Iter. # 1

10

15

Figure 7.9

50

100
150
200
Shear Wave Velocity [m/sec]

250

Shear Wave Velocity Profile from UEQMA Inversion Algorithm at Treasure


Island NGES

214

Experimental Results
8
7

Exact RMS Misfit

6
5
4
3
2
1
0
0

10

Iteration #

Figure 7.10

Convergence of UEQMA Inversion Algorithm at Treasure Island NGES

Figure 7.11 and 7.12 summarize the results obtained from the linear inversion of the
experimental attenuation curve. For the selected smoothing parameter, the RMS error
between the theoretical and the experimental attenuation curves is equal to 3.41.
0

0.05
Experimental

Attenuation Coefficient (1/m)

0.04

Depth [m]

10

15
0

Figure 7.11

1
2
Shear Damping Ratio [%]

Theoretical

0.03

0.02

0.01

0
0

20

40
Frequency (Hz)

60

80

Shear Damping Ratio Profile and Theoretical Attenuation Curve from


UEQMA Inversion Algorithm at Treasure Island NGES

Experimental Results

215
4

rms Error

3.5

3.41

2.5

2
-10
10

10

-5

10

10

10

10

Smoothing Parameter

Figure 7.12

Attenuation Curves RMS Misfit Error using UEQMA Inversion Algorithm


at Treasure Island NGES

Phase Velocity [m/sec]

250
Theoretical

200

Experimental
Iter. # 6

150
100

Iter. # 1

50
0
0

20

40
Frequency [Hz]

60

80

Attenuation Coefficient [1/m]

0.1

Figure 7.13

Theoretical

0.08

Experimental

Iter. # 1

0.06
0.04
Iter. # 6
0.02
0
0

20

40
Frequency [Hz]

60

80

Fundamental Mode Theoretical and Experimental Dispersion and


Attenuation Curves at Treasure Island NGES

216

Experimental Results

7.4 Coupled Inversion


This section will illustrate the results of the coupled inversion of the dispersion and
attenuation curves measured at Treasure Island NGES. They were obtained using the
fundamental-mode-based analysis of the CFUMA algorithm.
Figure 7.13 illustrates the convergence of the inversion algorithm in terms of dispersion
and attenuation curves. The theoretical curves corresponding to the sixth and final iteration
are in reasonable agreement with the experimental curves.
The corresponding sequences of shear wave velocity and shear damping ratio profiles
are shown in Fig.7.14 where the dashed lines indicate the starting models used in the
inversion. The final shear wave velocity and shear damping ratio profiles are shown using
bold lines.
0

Iter. # 6
Iter. # 6

Depth [m]

Depth [m]

10

10

Iter. # 1
Iter. # 1
15

15
0

50

100

150

200

250

Shear Wave Velocity [m/sec]

Figure 7.14

Shear Damping Ratio [%]

Shear Wave Velocity and Shear Damping Ratio Profile from CFUMA
Inversion Algorithm at Treasure Island NGES

Finally Fig.7.15 shows the convergence of the algorithm in terms of the RMS error
between the experimental and theoretical complex phase velocities; the convergence of the
algorithm after six iterations is apparent.

Experimental Results

217

Exact RMS Misfit

0
0

Figure 7.15

4
6
Iteration Num b e r

10

Convergence of CFUMA Inversion Algorithm at Treasure Island NGES

7.5 Results and Discussion


The results obtained from the uncoupled and coupled inversion of the experimental
dispersion and attenuation curves illustrated in Section 7.3 and Section 7.4 are now
compared with independent in-situ and laboratory measurements of shear wave velocity and
shear damping ratio. The cross-hole shear wave velocity data presented in Fig.7.16 are from
five different source-to-receiver travel paths between the boreholes shown in Fig. 7.1.
The in situ shear damping ratios shown in Fig.7.17 were obtained from analysis of crosshole seismic data utilizing a seismic waveform matching technique (Tang, 1992). They
represent the mean values and the associated uncertainties measured between three different
pairs of the boreholes shown in Fig.7.1.
Figures 7.16 and 7.17 show a comparison between shear wave velocity and shear
damping ratio profiles determined from the inversion of surface wave data, and the values
of these parameters obtained from independent in situ and laboratory measurements.
The values of shear wave velocity obtained from the interpretation of surface wave test
compare reasonably well with cross-hole measurements. Because Treasure Island NGES is a
relatively homogeneous site, there are small differences between the shear wave velocity
profiles predicted by the algorithms UFUMA, UEQMA, and CFUMA. As mentioned
several times in this dissertation, in regular soil profiles the response of the medium is
mainly controlled by the fundamental mode of propagation. It is therefore natural to expect
very similar results from the UFUMA, UEQMA, and CFUMA algorithms. From Fig.7.16

218

Experimental Results

the increasing lack of resolution with depth that is inherent to surface wave tests should also
be noted, particularly when using relatively small sources. This limitation can be partially
overcome by using more powerful sources capable of generating lower frequencies (< 5
Hz).

100
0

Shear Wave Velocity [m/sec]


125
150
175

200

CFUMA
UFUMA

Depth [m]

UEQMA
6

Cross-Hole

12

15

Figure 7.16

Comparison at Treasure Island NGES of Shear Wave Velocity from Surface


Wave Test Results with Other Independent Measurements

Experimental Results

219

Shear Damping Ratio


0%

2%

4%

6%

8%

10%

Depth [m]

CFUMA
UFUMA

12

UEQMA
Cross-Hole

15

Resonant C.
Torsional S.

Figure 7.17

Comparison at Treasure Island NGES of Shear Damping Ratio from


Surface Wave Test Results with Other Independent Measurements

Concerning the predictions of shear damping ratios, Fig.7.17 shows that the values
obtained from the interpretation of surface wave tests are generally less than those obtained
from cross-hole measurements. These differences can be attributed to three possible causes:
1. Frequencies used in cross-hole tests are usually on the order of several hundred Hertz or
more. The shorter wavelengths associated with these higher frequencies are more
susceptible to apparent attenuation due to scattering and other phenomena than are the
long wavelengths associated with surface wave testing at lower frequencies (i.e., 5 to 100
Hz).

220

Experimental Results

2. At higher frequencies involved in cross-hole tests, soil damping may be strongly affected
by fluid flow losses in addition to frictional losses as described earlier. These fluid losses
may cause the damping to increase and become frequency dependent at higher
frequencies.
3. There are substantial differences in the volume of soil sampled by the two methods.
The borehole spacing in the cross-hole measurements was approximately 3 m while the
surface wave measurements were performed with geophone offsets as large as 60 m.
Thus, the cross-hole measurements yield localized attenuation properties, and the
surface wave measurements yield properties which are averaged over a much larger
volume of soil. The extent to which the measurements differ will depend on the
heterogeneity of the site.
Laboratory-measured shear damping ratios obtained from resonant column and
torsional shear tests are also available for comparison at the Treasure Island NGES site.
Laboratory values of shear damping were measured on undisturbed soil specimens at
depths of 5.3 and 9.1 m from the borings drilled for the cross-hole tests. The laboratorymeasured values of shear damping are also shown in Fig.7.17. At 5.3 m, the agreement
between the laboratory and surface wave values of shear damping ratio is excellent. At 9.1
m, the laboratory values are slightly greater than those determined from surface wave tests.
For the resonant column tests, the damping ratios were measured at frequencies of
approximately 40 to 50 Hz. The torsional shear tests were performed at 0.5 Hz.

8 CONCLUSIONS AND RECOMMENDATIONS

8.1 Conclusions
Surface wave tests are non-invasive field techniques that can be used to determine the
low-strain dynamic properties of soil deposits. In this dissertation a different approach to
the conventional interpretation of surface waves measurements has been presented. The
new approach is developed around three fundamental ideas.
First, the definition of the so-called dynamic properties of soils is revisited within the
framework of a consistent theory of mechanical behavior. Often in the geotechnical
literature, the terms stiffness and damping ratio are used, in a loose sense, as if they were
intrinsic properties of soils. In reality, the definition of these terms is inherently linked to an
assumed constitutive model to which these behavioral properties are ascribed. Different
idealizations of material behavior will result in different types of behavioral properties.
Accordingly, although it is customary in geotechnical engineering to take for granted the
significance of terms like stiffness and damping ratio, it is important to reexamine the
definitions of these important mechanical parameters in the context of a consistent theory
of material behavior.
In Chapter 2 the definition of the low-strain dynamic properties of soils is formulated
within the framework of the linear theory of viscoelasticity. Experimental evidence
demonstrates that soils subjected to dynamic excitations at strain levels below the linear cyclic
threshold strain exhibit the ability to both store and to dissipate strain energy over a finite
period of time. At these low-strain levels, phenomena of instantaneous energy dissipation
are usually negligible. A measure of the energy dissipated within the soil mass during the
process of deformation is the internal entropy density production, which from experimental
observations appears to be a rate-independent quantity within the seismic frequency band.
These features of low-strain dynamic behavior of soils can accurately be modeled by the
theory of linear viscoelasticity.
The limitations and inconsistencies of two constitutive models commonly used in
geotechnical earthquake engineering and soil dynamics, the viscous Kelvin-Voigt model and
the rate-independent Kelvin-Voigt model, were noted. Both models fail to reproduce
fundamental features of soil behavior such as the instantaneous elastic response exhibited by
a soil specimen subjected to a suddenly applied stress and the important phenomenon of
stress relaxation. Furthermore, it was pointed out that the rate-independent Kelvin-Voigt
model violates both the time-translation invariance hypothesis of linear viscoelasticity, and
when applied to wave propagation problems, the most elementary principles of causality.

221

222

Conclusions and Recommendations

A consistent viscoelastic constitutive model requires the definition of a set of material


response functions that can be specified in either the time or the frequency domain. In this
context an important distinction has been made between the conventional Low-Strain
Dynamic Properties of Soils (LS-DPS), which are stiffness and material damping ratio, and the
Low-Strain Viscoelastic Properties of Soils (LS-VPS), which are the truly fundamental parameters
of a linear viscoelastic model. The relationships between these two sets of model parameters
have been established, and a new definition of material damping ratio has been introduced.
The necessity for a new definition of damping ratio is justified from the need to overcome
the limitations of the conventional definition when applied to strongly dissipative media.
For wave propagation problems it was found convenient to introduce another set of
viscoelastic parameters which are the body wave phase velocities and attenuation
coefficients collectively denoted as the Low-Strain Kinematical Properties of Soils (LS-KPS). The
three sets of model parameters, LS-VPS, LS-DPS, and LS-KPS, are equipollent; they simply
provide alternate ways to characterize the mechanical response of linear viscoelastic
materials.
One of the results obtained from reformulating the definitions of the low-strain
dynamic properties of soils is the recognition that stiffness and material damping ratio are
not independent parameters even in weakly dissipative media. The coupling between
stiffness and material damping ratio is the natural consequence of material dispersion, a
phenomenon that is formally defined by the Kramers-Krnig relations. Material dispersion
is an intrinsic feature of any type of viscoelastic materials. As a result of their mutual
dependence, a correct experimental procedure for determining stiffness and material
damping ratio of soils should determine these two parameters simultaneously. However, in the
current practice of geotechnical engineering testing, stiffness and material damping ratio are
determined separately using different measurement techniques. Based on these
considerations, the final part of Chapter 2 is dedicated to illustrating the principles of a new
experimental procedure to be conducted in laboratory with the resonant column test for the
simultaneous determination of soil stiffness and material damping ratio. The procedure is
also suitable for the investigation of the frequency dependence laws of these important soil
parameters.
After reexamining the definition of the low-strain dynamic properties of soils within the
framework of the theory of linear viscoelasticity, the focus of this dissertation shifted to
procedures used to determine these important soil parameters from the interpretation of
surface wave measurements. With regards to this subject, two new ideas were developed and
presented in this study. Together they form the basis of a new approach to surface wave
measurement and interpretation. The first contribution was the derivation of an explicit
representation of the effective Rayleigh phase velocity, which was defined as the phase
velocity of a superposition of a finite number of harmonic waves having the same frequency
and different wavenumbers. The effective Rayleigh phase velocity corresponds to the phase

Conclusions and Recommendations

223

velocity measured in surface wave testing with harmonic source if the near field effects are
neglected.
Surface waves propagating in vertically heterogeneous media are dispersive, and their
velocity of propagation is a multi-valued function of the frequency of excitation. This
phenomenon, known in the literature as geometric dispersion, arises from a condition of
constructing interference among rays that are either bent or reflected/refracted by the
heterogeneity of the medium. Geometric dispersion is responsible for the existence of
several modes of propagation each traveling at a different phase and group velocity in a
vertically heterogeneous medium. Geometric dispersion also affects the geometric spreading
of Rayleigh waves in homogeneous media.
Mathematically, the modes of propagation are obtained from the solution of a
differential eigenproblem, where the boundary conditions are 1) the vanishing of tractions
at the free surface of the half-space and 2) the radiation condition at infinity. The continuity
of stress and displacement fields must also be enforced in vertically heterogeneous media
where the material properties vary discontinuously with depth (i.e. multi-layered media). At
any given frequency of excitation, non-trivial solutions of this differential eigenproblem are
obtained only for special values of the wavenumber, which are called eigenvalues. In an elastic
medium, each eigenvalue is associated with a real-valued function called the eigenfunction that
gives the depth-variation of the displacement and stress fields for each mode of
propagation. Solution of the differential eigenproblem corresponds to the solution of the
free vibration or homogeneous problem of Rayleigh waves where no sources or initial
conditions are specified. A natural question then arises about the solution of the
inhomogeneous Rayleigh wave problem, particularly with regard to the velocity of propagation
of the ensuing surface wave field. This is relevant in surface wave testing because phase
velocity measurements are the basis for determining, via an appropriate inversion
procedure, the shear wave velocity profile at a site.
For Rayleigh waves generated by harmonic sources, the various modes of propagation
of surface waves are superimposed as in a spatial Fourier series. The phase velocity of the
resulting waveform was named effective Rayleigh phase velocity. In Chapter 3 an explicit
expression for the effective Rayleigh phase velocity has been derived. An important feature
of this kinematical quantity is its local nature, which makes its current value to be a function
of the spatial position where it is measured. If the contribution of the body wave field is
neglected (a valid assumption at distances of more than one to two wavelengths from the
source), then the effective phase velocity is the phase velocity that would be measured at a
point during a surface wave test. For measurements at two or more receiver locations, the
measured phase velocity is equal to the averaged effective phase velocity over the receiver
array. The notion of effective phase velocity and its explicit representation formed the basis
of a new interpretation of surface wave measurements where the modal dispersion curves
are replaced by the effective dispersion curve. The latter is obtained from the frequency
dependence law of the effective phase velocity averaged over the receiver array.

224

Conclusions and Recommendations

This new approach overcomes a major problem affecting the conventional


interpretation of surface wave testing: the inconsistency of matching an experimental
dispersion curve that reflects, in general, the contributions of several modes of propagation
with a simulated dispersion curve for only a single mode of propagation. The current
strategy used by several researchers to overcome this inconsistency is to perform a
numerical simulation of the actual experiment. With this approach the theoretical phase
velocities are computed from phase differences of theoretical displacements that are
calculated at locations that emulate those used in the actual SASW test. Although this
method is exact, it has the disadvantage of requiring the computation of the displacement
field, which is computationally more expensive than merely solving the Rayleigh
eigenproblem if the body wave field is included. Another disadvantage of this method is
that the partial derivatives required for the solution of the non-linear inverse problem have
to be computed numerically, a task that is computationally more expensive than using
analytical partial derivatives and also potentially inaccurate.
In the new approach to surface wave interpretation, the experimental dispersion curve is
compared and matched with the effective dispersion curve, which accounts for multi-mode
Rayleigh wave propagation. In Chapter 3 closed-form analytical expressions for the partial
derivatives of the effective Rayleigh phase velocity with respect to the shear and
compression wave velocities of the layers have been derived by employing the variational
principle of Rayleigh waves. The major advantage offered by the analytical over the
numerical partial derivatives, is that the former are computed using the solution of the
Rayleigh eigenproblem referred to the original and not the perturbed profile of medium
parameters.
The second contribution in the area of surface wave propagation presented in this study
is a numerical technique for the solution of the Rayleigh eigenvalue problem in linear
viscoelastic media. One immediate application of this result was the development of a
systematic and efficient procedure for simultaneously determining the low-strain dynamic
properties of soil deposits from the interpretation of surface wave measurements. In the
conventional interpretation of surface wave data, the shear wave velocity and the shear
damping ratio profiles at a site are determined separately from the inversion of an
experimental dispersion and attenuation curve. The simultaneous inversion of surface wave
data offers several advantages over the corresponding uncoupled analysis.
First, it explicitly recognizes and accounts for the inherent coupling existing between
seismic wave phase velocity (which is directly related to stiffness) and material damping ratio
as a consequence of material dispersion. Secondly, the simultaneous inversion is a better-posed
mathematical problem (in the sense of Hadamard). The solution of the (strongly) coupled
Rayleigh inverse problem is based on the use of a complex formalism where the Rayleigh
phase velocity is viewed as an holomorphic function VR* ( VS* ) of the complex-valued shear
wave velocity. Thus, the simultaneous inversion takes full advantage of the internal constraint

Conclusions and Recommendations

225

constituted by the Cauchy-Riemann equations satisfied by the analytic function VR* ( VS* ) .
Thirdly, the simultaneous inversion eliminates some of the errors affecting the
corresponding uncoupled analysis, where some of the input data required for the inversion
of the attenuation measurements are obtained from the inversion of the dispersion data,
and thus they are affected by the uncertainties associated with the latter process. In the
simultaneous inversion, this problem is overcome by inverting both dispersion and
attenuation measurements in a single, complex-valued, inversion procedure.
It should be noted that the simultaneous inversion presented in this study is
conceptually different from other types of simultaneous inversions available in the literature.
A fully coupled Rayleigh inversion requires the ability to solve the complex eigenproblem in
linear viscoelastic media, where Rayleigh phase velocity and attenuation depend upon both
the body wave velocities and material damping ratios of the medium. This type of inversion,
here denoted as strongly coupled, has been implemented in this study by introducing a new
technique for the solution of the complex Rayleigh eigenproblem.
In general, most of the difficulties associated with the solution of an eigenproblem
(differential or algebraic) are related with the determination of the eigenvalues. In the case
of the complex Rayleigh eigenproblem, this task involves computing the roots of the
complex-valued Rayleigh dispersion equation. This is not a trivial problem, particularly
because the Rayleigh dispersion equation is highly non-linear and is known only numerically.
The technique used in this work to accomplish this task is an elegant procedure based on
the use of Cauchy residue theorem of complex variable theory. Once the roots of the
dispersion equation are calculated, the correspondence principle of linear viscoelasticity is
invoked for the computation of the eigenfunctions, the effective Rayleigh phase velocity,
the modal and the effective partial derivatives of Rayleigh phase velocity with respect to the
complex-valued body wave velocities of the medium.
Often in the seismological literature the term simultaneous inversion is used to denote
an approximate procedure that, based on the assumption of weak dissipation, does not
actually require the solution of the complex eigenproblem. Conversely, it uses a variational
approach combined with the results obtained from the solution of the elastic eigenproblem
to find an approximate solution to the complex eigenproblem. This type of approach is
here denoted as the weakly coupled inversion. The technique presented in this study for the
solution of the complex eigenvalue problem is not restricted to weakly dissipative media.
The theory of modal and effective Rayleigh waves propagation in elastic and viscoelastic
media developed in Chapter 3 has been used in Chapter 4 to develop four types of
inversion algorithms named UFUMA, UEQMA, CFUMA, and CEQMA. For a given pair
of experimental dispersion and attenuation curves, these algorithms determine the shear
wave velocity and shear damping ratio profiles of a soil deposit. The inversion procedure
used in these four algorithms is based on a constrained least squares algorithm known as
Occams algorithm. Its main objective is to enforce maximum smoothness on the resulting shear

226

Conclusions and Recommendations

wave velocity and shear damping ratio profiles while attaining a specified error misfit
between the experimental and the simulated dispersion and attenuation curves. The choice
of this inversion strategy was motivated by the need of minimizing the dependence of the
inverted shear wave velocity and shear damping ratio profiles upon the assumed number of
layers, which is, in general, an additional unknown in the interpretation of surface wave
testing.
The algorithms UFUMA and UEQMA refer to the uncoupled fundamental mode and
uncoupled equivalent multi-mode inversion analysis, whereas the algorithms CFUMA and
CEQMA were designed to perform the coupled fundamental mode and coupled equivalent
multi-mode inversion analysis, respectively.
In Chapter 6 these algorithms were tested in a systematic numerical simulation involving
a homogeneous medium and three simplified, stratified media. Case 1 was a regular soil
profile where the stiffness increases regularly with depth, whereas Case 2 and Case 3 were
two different types of irregular soil profiles. In this numerical simulation, the synthetic
dispersion and attenuation curves were determined using the same procedures used in an
actual SASW experiment and illustrated in Chapter 5. From the numerical simulation, the
results obtained with the algorithm CEQMA were considered unreliable due to an inability
displayed by this algorithm to determine the correct sequence of Rayleigh modes at certain
frequencies.
Concerning the results obtained with the inversion algorithms UFUMA, UEQMA, and
CFUMA, they can be summarized as follows:
1. Overall, the UEQMA inversion algorithm based on the concept of effective phase
velocity yielded the most accurate results, particularly for the prediction of the shear
wave velocity profile. The fundamental mode based algorithms UFUMA and CFUMA
were generally less accurate in predicting both the shear wave velocity and shear
damping ratio profile;
2. For regular soil profiles (Case 1) the fundamental mode based inversion yielded
satisfactory results. However the results obtained with the coupled inversion algorithm
CFUMA are more accurate than those obtained with the corresponding uncoupled
analysis performed with the algorithm UFUMA;
3. For irregular soil profiles (Case 2 and Case 3) the results of the algorithm UEQMA were
always more accurate than those obtained with the algorithm UFUMA and CFUMA.
The results obtained from the numerical simulation are consistent with the expectations
from the theory, and also with the results of other independent studies (Gucunski and
Woods, 1991; Tokimatsu, 1995). Whereas in regular soil profiles the fundamental mode of
propagation governs the response of a layered medium to a dynamic excitation, irregular soil

Conclusions and Recommendations

227

profiles are inversely dispersive, and as such the response of the medium includes
contributions from higher modes of propagation.
8.2 Recommendations for Future Research
Surface wave propagation with the associated inverse problem are fascinating and
challenging subjects whose need for further investigation is justified by reasons that extend
far beyond the practical applications. Areas of research where some of the ideas presented
in this study may be further expanded include:
1. Implementation of a routine for the subdivision of the region of the complex plane
containing the roots of the Rayleigh secular function into a series of smaller subregions.
As explained in Section 3.6.1, the computation of the roots of the Rayleigh secular
function via a high-degree polynomial becomes an ill-conditioned problem as the
number of modes increases. As a result these roots are computed with a decreasing
degree of accuracy. The instabilities exhibited by the code CEQMA during the
computation of higher modes of propagation are most likely due to this problem. The
technique of subdividing the region containing the roots should stabilize the
performance of the algorithm CEQMA.
2. Implementation of a new method for the construction of the Rayleigh secular function
in linear viscoelastic media to be integrated with the root-finding technique presented in
this study. The actual technique for constructing the Rayleigh secular function is based
on the method of reflection and transmission coefficients. This method, originally
developed for elastic media, does not seem to be very accurate when applied to
viscoelastic systems, particularly at high frequencies and for large number of layers. The
spectral element method, the boundary element method and numerical integration are
possible alternative methods.
3. In surface wave tests the quantities measured experimentally are displacement spectra or
displacement transfer functions. In this study it was shown that Rayleigh phase velocity
and attenuation coefficient are derived quantities that are obtained from the displacement
spectra or transfer functions via an unstable process of numerical differentiation. Based
on this observation it would be interesting to attempt the construction of an algorithm
for determining the medium parameters from the direct inversion of the displacement
spectra or transfer functions.
4. Procedures should be developed for solving the Rayleigh inverse problem by adopting
Global-Search-Techniques such as genetic algorithms, fractal inversion, neural network
inversion, or Monte Carlo simulation. As mentioned in Chapter 4, these methods are
more robust and accurate than Local-Search-Techniques such as Occams algorithm,
even though they are computationally more expensive.

228

Conclusions and Recommendations

5. Investigate if it is advantageous to re-cast the current interpretation of surface wave


testing using a wavelets-based analysis. From a theoretical point of view investigate the
possibility of expanding the solution of the Rayleigh eigenproblem using a wavelets
multi-resolution analysis.
6. Solution of the boundary value problem of surface waves using more sophisticated
constitutive laws other than classical linear viscoelasticity. Examples include binary
porous media theories, non-local and polar theories, doublet-mechanics. Attempt to
relax the usual assumptions of small displacements/displacement gradients.

APPENDIX A - ELLIPTIC HYSTERETIC LOOP IN LINEAR


VISCOELASTIC MATERIALS

A.1 Harmonic Constitutive Relations


The stress-strain relationships of linear viscoelastic materials subjected to harmonic
excitations assume a particularly simple form:

( ) = G * ( ) ( )

(A.1)

where ( ) = 0 e it , 0 R , G * ( ) is the complex modulus, and = P, S is a


subscript denoting the irrotational and the equivoluminal (shear) components of the
associated tensorial quantity. By considering the real part of ( ) , Eq. A.1 can be rewritten
as:

( ) = G* ( ) 0 cos t ( )

(A.2)

where tan ( ) = arg G * ( ) is the loss angle. Using trigonometric identities Eq. A.2 can
be rewritten as:

( ) G* ( ) ( ) cos ( ) = G* ( ) 0 sin ( ) sin ( t)

(A.3)

Equation A.3 combined with the relationship ( ) = 0 cos( t ) gives:


2

2
G(1)

=1

( 2) 0
0

(A.4)

229

230

Appendix A

where G(1) and G( 2 ) are the real and the imaginary parts of the complex modulus G * ( ) .
Equation A.4 is the equation of an ellipse rotated by an angle ( ) with respect to the
strain axis (see Fig. 2.9). It represents the stress-strain hysteretic loop exhibited by a linear
viscoelastic material subjected to harmonic oscillations.
A.2 Energy Dissipated in Harmonic Excitations
The area enclosed by the elliptic hysteretic loop can be interpreted as the amount of
energy (per unit volume) dissipated by the material during a cycle of harmonic loading. In a
dissip
stress-controlled test, this area, here denoted by W () , is defined as:

( ) ( )

Wdissip ( ) = d W = d
l

(A.5)

where the symbol () denotes the real part of a complex quantity, and l is length of the
hysteretic loop. In Eq. A.5 the term dW represents the work done by the stress (per unit
volume of the material) for an infinitesimal variation of the strain. Considering Eq. A.2 and

( )

the fact that d = 0 cos (t + 2) dt , Eq. A.5 can be rewritten as:

Wdissip ( ) =

G * 20 cos t cos (t + 2) dt

(A.6)

Using trigonometric identities this integral can be simplified to:

( )

Wdissip ( ) = G * 20 sin

sin 2 ( t) dt

(A.7)

which can be easily solved to give:

Wdissip ( ) = G( 2)

(A.8)

Appendix A

231

A.3 Principal Axes of the Elliptic Hysteretic Loop


To obtain the inclination of the ellipse principal axes, it is convenient to rewrite Eq. A.4
as follows:

G 2
G
2

(
)
(
)
1
1
2
1

2

1+
2 +
2 =1
20 G( 2 )

G( 2 )
G

0 ( 2 )

(A.9)

Then, utilizing the result that the coefficients of a quadric centered at the origin and written in
the form A ik x i x k = 1 represent the components of a second order tensor (Finzi and Pastori,
1961), it is easily recognized from Eq. A.9 that:

G 2
G(1)
(1)
1

2
2 1 + G( 2 )
G( 2 )

0
0

A=

G(1)

1

2
2
G
G

0
(2)

0 (2)

(A.10)

The problem of finding the ellipse principal axes has been transformed in a problem of linear
algebra, namely of finding the eigenvectors of the matrix A. Because this matrix is real and
symmetric, its eigenvalues are also real and the corresponding eigenvectors, as expected, are
orthogonal. A straightforward computation yield for the eigenvalues 1,2 :

1 ,2 ( ) =

G * 2 + 1 G * 2 + 1 4G 2
( 2 )

2 0 G( 2 )

(A.11)

The inclination of the ellipse principal axes 1 = ( ) and 2 = 2 +


from the components of the eigenvectors; the result is:

) is obtained

232

Appendix A
2

tan 1,2 ( ) =

G* 2 1 m G* 2 + 1 4G 2
( 2)


2 G(1)

(A.12)

APPENDIX B - EFFECTIVE RAYLEIGH PHASE


VELOCITY PARTIAL DERIVATIVES
In Section 3.5.2 it was shown that the first variation of the effective Rayleigh
phase velocity V$ ( = r , y ) can be written as:
V$

V$
V$ =
Vj +
U j
U j
Vj

(B.1)

where the summation convention is implied over the index j ranges from one to the
number of Rayleigh modes of propagation associated with the frequency (cf.
Eq.3.49). Explicit results will now be obtained for the terms Vj and U j of Eq. B.1.
From the relations G = VS2 , ( + 2G) = VP2 and Eq. 3.41, the following result is
obtained:

Vj = PjVS + Q jVP dy
0

(B.2)

where:
Pj (y , ) =

Q j ( y , ) =

VS

2 ( k 2 U I1 ) j

VP

dr1
dr
4 k r1 2
kr2
dy
dy

2 ( k 2 U I1 ) j

(B.3a)

dr
kr1 + 2
dy j

(B.3b)

The strategy to obtain an explicit relation for U j is a bit more laborious. Application
of Rayleigh principle to Eq.3.33a gives I1 = 0 , hence Eq.3.50 simplifies as follows:

233

234

Appendix B

I
I2
I
U j = 2 2 V +
3
V I1 2I1 j
V I1

(B.4)

The first variations I 2 and I 3 can be calculated from Eq.3.33 considering again
Rayleigh principle and the relations G = VS2 and ( + 2G) = VP2 ; the result is:

I 2 = ( VP r12 VP + VS r22 VS )dy

(B.5a)

dr
dr
dr
I 3 = 2 VP r1 2 VP VS r2 1 + 2 r1 2 VS dy
dy
dy
dy
0

(B.5b)

Substitution of Eq.B.5 and Eq.B.2 in Eq.B.4 yields:

U j =

[ V
j

+ jVP dy

(B.6)

where:
j ( y , ) =

j ( y , ) =

2 (I1 ) j
1

2 (I1 ) j

dr2
dr1
( k 2 I2 ) j Pj
VS kr22 r2
2r1
dy j
dy

(B.7a)

dr
VP kr12 r1 2 ( k 2 I2 ) j Q j
dy j

(B.7b)

In light of Eq.B.2 and Eq.B.6, Eq.B.1 can be rewritten as follows:

V$
V$
V$
V$
$
V =
Pj +
j VS +
Qj +
j VP dy
U j
U j
Vj

0
Vj

which suggests the following result (cf. Eq.3.42):

(B.8)

Appendix B

235

$
V =
0

V$
VS dy +

VS , VP

V$

VP dy

VP ,VS

(B.9)

where:
V$
V$
V$

Pj +
=

U j j
VS ,VP Vj
V$
V$
V$

=
+

U j j
VP , VS Vj
Equation B.10 provides an explicit relationship for the partial derivatives of the
effective Rayleigh phase velocity V$ (r , y , ) with respect to the medium parameters
VP and VS . A complete definition of this relation however, requires specification of
V and V
U , whose calculation will be the objective of the rest
the terms V

of the Appendix.
If in Eq.3.47 the modal group velocity U k ( k = i , j) is held constant, Eq.3.46 can
be differentiated (in the Gateaux sense) with respect to Vk , to yield:

(B.10)

236

Appendix B

[V$ (r, y, )]

V Vj
1
C ijr 2i + 2 Vi Vj Dij Vi Vj
ij
ij
Vj
2
Vi
2 M M

A i=1 j=1
Vi Vj

1
2

( ) ( ) (V V + V V )

( )

V Vj
1
1 V Vj
Dij 2i 2 r C ij + 2i + 2 Vi Vj
ij
ij
Vj
Vj
Vi Vj Vi
2B M M
Vi
+

Vi Vj
A 2 i=1 j=1

( )

( )

1
1 1
Dij + Vi Vj
ij
Vi Vj 2
2B M M

Vi Vj
A 2 i=1 j=1

1
2

( ) (V V + V V )

( )

(B.11)
where:

( ) (1 V + 1 V ) cos [r (1 V 1 V )]



A (r , y , ) =
i =1 j =1

( )



B (r , y , ) =
i =1 j =1

ij

Vi Vj

ij

[ (

)]

cos r 1 Vi 1 Vj

V i Vj

(B.12a)

(B.12b)

and:

[ (

)]

[ (

)]

C ij (r , ) = sin r 1 Vi 1 Vj

D ij (r , ) = cos r 1 Vi 1 Vj

(B.13a)
(B.13b)

Appendix B

237

After some algebra, Eq.B.11 can be reduced to the following form:

[V$ (r , y, )]

2 M M
A
A 2 i=1 j=1

( )

( )

( )

B Vi A
ij
ij


B Vj
ij
ij

( )

(B.14)

where:

( ) (2rC
( ) =
2( V V )

ij

ij

( )

ij

ij

ij

ij

ij

ij
2

ij

)]

Vi Vj

+ Vj D ij Vi2

( ) [2(rV C

Vi Vj D ij + rVi C ij Vi D ij Vi + Vj Vj2
2 Vi Vj

ij

(B.15a)

V i Vj

( ) [2(rV C

( )

( ) (2rC
( ) =
2( V V )

Vi D ij Vj2

ij

(B.15c)

Vi Vj

ij

)]

+ Vi VjD ij + rVi C ij + VjD ij Vi + Vj Vi2

( )

2 Vi Vj

(B.15b)

Vi Vj

(B.15d)

Equation B.14 can be re-written in a more compact form by further setting:

( )

ij

( )

2 A

( )

B
ij
A 2
ij

(B.16)

( )
~

ij

( )

2 A

( )

B
ij
ij
2
A

238

Appendix B

so that:

V$ (r , y , )

( )

( )

~
= Vi Vj
ij
ij

i =1 j =1

(B.17)

By replacing Eq.B.2 for the terms Vi and Vj in Eq.B.17, the latter becomes:

V$ (r , y , )

M M
~

= Pi Pj VS dy +
ij
ij

0
i =1 j =1

( )

( )

(B.18)
M M

Q
Q

VP dy
ij
ij

i =1 j=1

( )

( )

Now the next task is the computation of V$ (r , y , ) . At this purpose Eq.3.46


V

is differentiated with respect to U k while keeping the modal phase velocity Vk


constant, yielding:

[V$ (r , y , )]

U U j

D ij i +
ij
U j
Ui
2 M M

A i =1 j =1
Vi Vj

( )

1
1 U U j

D ij + i +
ij
U j
Vi Vj U i
2B M M

A 2 i=1 j=1
Vi Vj

( )

Equation B.19 can be reduced, after some algebra, to a form analogous to that of
Eq.B.17, namely:

(B.19)

Appendix B

239

V$ (r , y , )

( )

( )

~
= T U i + T U j
ij
ij

i =1 j =1

(B.20)

where:

(T )

(~T )

( )

= U j W

ij

( )

= U i W

ij

ij

(B.21)
ij

and:

(W )

( )

ij

ij D ij
2 1
1
= 2 B + A
A Vi Vj
U i U j Vi Vj

(B.22)

If the terms U i and U j in Eq.B.20, are replaced by Eq.B.6, the former


becomes:

V$ (r , y , )

M M

~
= T i + T j VS dy +
ij
ij

i =1 j=1
0

( )

( )

(B.23)
M M

T
T

0 i =1 j=1 ij i ij j VP dy

( )

( )

Combining the results of Eq.B.18 and Eq.B.23 in light of Eq.B.1:

240

Appendix B

[V$ (r , y , )] = [V$ (r , y , )] + [V$ (r , y , )]

M M


~
~

+
+

P
P
T
T

0 i=1 j=1 ij i ij j ij i ij j VS dy +

( )

( )

( )

( )

(B.24)

M M

ij Q i ~ ij Q j + T ij i + ~T ij j VP dy

i =1 j =1
0

( )

( )

( )

( )

Finally, comparison of Eq.B.24 and Eq.B.9 yields:


M M
V$
V$

~
~
r , y , ) = Pi Pj + T i + T j
=

(
ij
ij
ij
ij

i =1 j=1
VS ,VP VS

( )

( )

( )

( )

M M
V$
V$

~
r , y , ) = Q i Q j + T i + T j
=

(
ij
ij
ij
ij

i =1 j =1
VP , VS VP

( )

( )

( )

( )

(B.25)

APPENDIX C - DESCRIPTION OF COMPUTER CODES

C.1

UFUMA (Uncoupled-Fundamental-Mode-Analysis)

Dispersion is the first module of the algorithm UFUMA used to perform the
Fundamental-Mode-Uncoupled-Inversion of experimental surface wave data at a site. In
particular, the module Dispersion is a MATLAB m-file designed to implement the non-linear
inversion of the experimental dispersion curve. The non-linear inversion is performed by using
a constrained-least-squares-algorithm called the Occams algorithm, which enforces maximum
smoothness to the resulting shear wave velocity profile of the site, while attaining a specified
error misfit between the experimental and the simulated dispersion curves.
The shear damping ratio profile is calculated independently by a constrained-linear
inversion of the experimental attenuation data also based on the Occams algorithm. This task is
implemented by the MATLAB m-file Damping, which constitutes the second module of the
algorithm UFUMA.
Dispersion uses a mex-file called Rayleigh which is a FORTRAN 77 written routine that
solves the eigenvalue problem of surface Rayleigh waves in elastic vertically heterogeneous
media (modified from Hisada, 1995). This code computes also the modal partial derivatives of
Rayleigh phase velocity with respect to the shear and compression wave velocities of the soil
layers using a variational formulation and the G-matrix formed by these partial derivatives. The
G-matrix and the Greens function associated with the Rayleigh wave-displacement field,
which is also computed by Rayleigh, are used by the module Damping for the inversion of
the experimental attenuation measurements.
The theoretical dispersion and attenuation curves computed by Dispersion and
Damping, as well as the modal partial derivatives of Rayleigh phase velocity with respect to
medium parameters, are calculated with respect to the fundamental mode of propagation of
Rayleigh waves.
All INPUT data required by the mex-file Rayleigh are imported via a data file from
Dispersion. The experimental phase velocities and attenuation coefficients are imported into
Dispersion and Damping from external ASCII files.

EXAMPLE DATA FILE


PARAMETERS USED IN THE NON-LINEAR INVERSION ALGORITHM

241

242

Appendix C

itmax = 10;
Maximum Number of Iterations Allowed
imumax = 2;
Maximum Number of Sub-Iterations Allowed
tolVS = 1e-2; Termination Criterion for Convergence
SITE INVESTIGATED: Treasure Island
Number of Layers
NL = 9;
Thickness of Layers (the half-space is denoted by 0.0)
THK = [1.52 1.52 1.98 2.44 3.05 9.14 10.00 10.00
Mass Density of Layers
DNS = [1.75 1.75 1.75

1.75

1.75

1.75

1.75

1.75

0.0]';

1.75]';

Compression Damping Ratio of Layers (need DP/DS in forming the G-matrix)


DP = [1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0]';
Elastic Poisson's Ratio of Layers
NU = [0.25 0.30 0.30 0.30 0.30

0.30

0.30

0.30

0.30]';

Shear Damping Ratio of Layers (need DP/DS in forming the G-matrix)


DS = [1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0]';
Initial Guess Shear Wave Velocity of Layers
VS0 = [60.0 60.0 60.0 60.0 60.0 60.0 60.0

60.0

60.0]';

FREQUENCY DATA (Hz)


Type of Frequency Spacing
IFREQ = 1;

[1 = logarithmic]

[0 = linear arithmetic]

Number of Frequencies
NF = 83;
Initial Frequency
IOM = 8.453;
Final Frequency
NOM = 70.823;
Frequency Increment (only for linearly spaced frequencies)
DOM = 0.0;
SOURCE PARAMETERS (SPATIAL POSITION AND MAGNITUDE)
Depth of the Source
DPH = 0.0;
Magnitude: X-Component
FXX = 0.0;
Magnitude: Y-Component
FYY = 0.0;
Magnitude: Z-Component
FZZ = 1.0;
RECEIVER PARAMETERS (Spatial Position in Cylindrical Coordinates)
Azimuth angle (Degrees)

Appendix C

243

PHI = 0.0;
Depth of the receivers
ZH = 0.0;
Number of Receivers
NP = 100;
Position First Receiver
INP = 1.00;
Position Last Receiver
NNP = 100.00;
Increment of Receiver Position
DNP = 1.00;
PARAMETERS USED TO COMPUTE ROOTS OF RAYLEIGH SECULAR FUNCTION
Specified Tolerance (TOL < 0.1)
TOL = 0.001;
Number of Partitions to Search Roots from CMIN to CMAX
NCC = 400;
Number of Sub-Partitions for Roots Close to VP and VS
NSC = 10;
Range of Velocity to Use Sub-Partitions
DCR = 20.0;
IMPORT EXPERIMENTAL SURFACE WAVE MEASUREMENTS
load ExpPhase.dat
load ExpAlpha.dat
END OF DATA FILE

C.2

UEQMA (Uncoupled-Equivalent-Multi-Mode-Analysis)

Dispersion is the first module of the algorithm UEQMA used to perform the EquivalentMulti-Mode-Uncoupled-Inversion of experimental surface wave data at a site. In particular, the
module Dispersion is a MATLAB m-file designed to implement the non-linear inversion of
the experimental dispersion curve. The non-linear inversion is performed by using a
constrained-least-squares-algorithm called Occams algorithm, which enforces maximum
smoothness to the resulting shear wave velocity profile of the site, while attaining a specified
error misfit between the experimental and the simulated dispersion curves.
The shear damping ratio profile is obtained independently by a linear inversion of the
experimental attenuation data also based on the Occams algorithm. This task is implemented by
the MATLAB m-file Damping, which constitutes the second module of the algorithm
UEQMA.
Dispersion uses a mex-file called Rayleigh which is a FORTRAN 77 written routine that
solves the eigenvalue problem of surface Rayleigh waves in elastic vertically heterogeneous

244

Appendix C

media (modified from Hisada, 1995). This code also computes the effective Rayleigh phase
velocity and its partial derivatives with respect to the shear and compression wave velocities of
the soil layers using a variational formulation. The effective G-matrix formed by these partial
derivatives is also computed. The effective G-matrix and the Greens function associated with
the Rayleigh wave-displacement field, which are also computed by Rayleigh, are used by the
module Damping for the inversion of the experimental attenuation measurements.
The effective theoretical dispersion and attenuation curves computed by Dispersion and
Damping are calculated by taking into account all the modes of propagation of Rayleigh
waves. Averaging the effective Rayleigh phase velocities and attenuation coefficients over the
frequency dependent receiver offsets eliminates the dependence of these quantities on the
receivers location.
All INPUT data required by the mex-file Rayleigh are imported via a data file from
Dispersion, which is the same data file used by the code UFUMA. The experimental phase
velocities and attenuation coefficients are imported into Dispersion and Damping via
external ASCII files.
C.3

CFUMA (Coupled-Fundamental-Mode-Analysis)

ViscoRay is the module of the algorithm CFUMA used to perform the FundamentalMode-Coupled-Inversion of the experimental dispersion and attenuation curves at a site. The
non-linear simultaneous inversion is performed by applying the complex formalism to a
constrained-least-squares-algorithm, called Occams algorithm, which enforces maximum
smoothness to the resulting complex shear wave velocity profile of the site, while attaining a
specified error misfit between the complex-valued experimental and simulated dispersion
curves.
ViscoRay is a MATLAB computer code interfaced with a mex-file called Rayleigh which
is a FORTRAN 77 routine that solves the complex eigenvalue problem of Rayleigh waves in
linear viscoelastic vertically heterogeneous media. This code also computes the partial
derivatives of the (modal) complex Rayleigh phase velocity with respect to the complex shear
and compression wave velocities of the soil layers using a variational formulation.
The theoretical dispersion and attenuation curves computed by the program ViscoRay, as
well as the modal partial derivatives of Rayleigh phase velocity with respect to medium
parameters, are referred to the fundamental mode of propagation of Rayleigh waves. However,
the frequency dependent attenuation coefficients are calculated using a geometric spreading
function that accounts for all the modes of propagation (Rix et al., 1998a).
All INPUT data required by the mex-file Rayleigh are imported via a data file from
ViscoRay, which is the same data file used by the codes UFUMA and UEQMA. The

Appendix C

245

experimental phase velocities and attenuation coefficients are imported into ViscoRay via
external ASCII files.
PARAMETERS USED TO COMPUTE THE ZEROS OF THE RAYLEIGH SECULAR FUNCTION
Number of Points in the Complex Plane used for Contour Integration with
the Gauss-Legendre Quadrature Formulae
Number of Points along curves DR = constant
NCC = 50;
Number of Points along curves CR = constant
NDD = 50;

C.4

CEQMA (Coupled-Equivalent-Multi-Mode-Analysis)

ViscoRay is the module of the algorithm CEQMA used to perform the Equivalent-MultiMode-Coupled-Inversion of the experimental dispersion and attenuation curves at a site. The
non-linear simultaneous inversion is performed by applying the complex formalism to a
constrained-least-squares-algorithm called Occams algorithm which enforces maximum
smoothness to the resulting complex shear wave velocity profile of the site, while attaining a
specified error misfit between the complex-valued experimental and simulated dispersion
curves.
ViscoRay is a MATLAB computer code interfaced with a mex-file called Rayleigh, which
solves the complex eigenvalue problem of Rayleigh waves in viscoelastic vertically
heterogeneous media. This code also computes the effective complex Rayleigh phase velocity
and its partial derivatives with respect to the complex shear and compression wave velocities
of the soil layers using a variational formulation.
The effective theoretical dispersion and attenuation curves computed by ViscoRay as well
as the effective partial derivatives of Rayleigh phase velocity with respect to medium
parameters, are calculated by taking into account all the modes of propagation of Rayleigh
waves. Averaging the effective Rayleigh phase velocities and the effective partial derivatives
over the frequency dependent receiver offsets eliminates the dependence of these quantities on
the receivers location. The experimental attenuation coefficients are calculated iteratively using
a geometric spreading function that also accounts for all the modes of propagation of Rayleigh
waves (Rix et al., 1998a).
All INPUT data required by the mex-file Rayleigh are imported via a data file from
ViscoRay, which is the same data file used by the code CFUMA including the parameters
used to compute the zeros of the Rayleigh secular function.

246

Appendix C

BIBLIOGRAPHY
Abd-Elall, L.F., Delves, L.M., Reid, J.K. (1970). A Numerical Method for Locating the
Zeros and Poles of a Meromorphic Function., from Numerical Methods for Non Linear
Algebraic Equations, by P. Rabinowitz, Ed. Gordon and Breach Science Publishers, pp. 4759.
Abo-Zena, A.M. (1979). Dispersion Function Computations for Unlimited Frequency
Values., Geophys. J. R. Astr. Soc., 58, 91-105.
Achenbach, J.D. (1984). Wave Propagation in Elastic Solids., North-Holland, Amsterdam,
Netherlands, pp. 425.
Aki, K., and Richards, P.G. (1980). Quantitative Seismology: Theory and Methods., W.H.
Freeman and Company, San Francisco, 932 pp.
Anandarajah, A. (1996). Discrete Element Method for Platy Colloidal Particles., Symposium,
Computational and Experimental Methods for Particulate Materials, ASME, Johns Hopkins
University, Baltimore, MD, June 12-14, 1996.
Anderson, D.L., and Archambeau, C.B. (1964). The Anelasticity of the Earth.,J. Geophysical
Research, 69(10), 2071-2084.
Anderson, D.L., Ben-Menahem, A., and Archambeau, C.B. (1965). Attenuation of Seismic
Energy in the Upper Mantle. J. Geophysical Research, 70, 1441-1448.
Azimi, S.A., Kalinin, A.V., Kalinin, V.V., and Pivovarov, B.L. (1968). Impulse and
Transient Characteristics of Media with Linear and Quadratic Absorption Laws.,
Izvestiya, Physics of the Solid Earth, February 1968, pp.88-93.
Baran, P.A., and Sweezy, P.M. (1968). Monopoly Capital: An Essay on the American
Economic and Social Order (Harmondsworth: Penguin Books).
Bth M. (1968). Mathematical Aspects of Seismology., Elsevier Publishing Company,
Amsterdam, pp.415.
Bellotti, R., Ghionna, V.N., Jamiolkowski, M., and Robertson, P.K. (1989). Design
Parameters of Cohesionless Soils from In-Situ Tests., Specialty Session on In-Situ Testing of
Soil Properties for Transportation Facilities, Sponsored by Committee A2L02-Soil and Rock
Properties, National Research Council, Transportation Research Board, Washington, January,
1989.

247

248

Bibliography

Bendat, J. and Piersol, A. (1986). Random Data - Analysis and Measurement Procedures.,
2nd Ed., John Wiley & Sons, New York, 566 pp.
Ben-Menahem, A., and Singh, S.J. (1981). Seismic Waves and Sources., Springer-Verlag,
New York, 1108 pp.
Biot, M.A. (1955). Theory of Elasticity and Consolidation for a Porous Anisotropic Solid.
Journal of Applied Physics, 26, 182-185.
Biot, M.A. (1956). Theory of Propagation of Elastic Waves in a Fluid-Saturated Porous
Solid., I. Lower Frequency Range; II. Higher Frequency Range, J. Acoust. Soc. Am., 28,
168-178; 179-191.
Boore, D.M. (1972). Finite Difference Methods for Seismic Wave Propagation in
Heterogeneous Materials. Methods of Computational Physics, Vol.11, Ed. Bolt, B.A.,
Academic press, New York, pp.1-36.
Bowen, R.M. (1982). Compressible Porous Media Models by Use of the Theory of
Mixtures. International Journal of Engineering Science, Vol. 20, No. 6, pp. 697-735.
Bracewell, R. (1965). The Fourier Transform and its Applications., McGraw-Hill Co.
Casagrande, A. (1932). Research on the Atterberg Limits of Soils., Public Roads 13 (8),
121-130 and 136.
Chen, X. (1993). A Systematic and Efficient Method of Computing Normal Modes for
Multilayered Half Space. Geophysics J. Int., Vol. 115, pp. 391-409.
Christensen, R.M. (1971). Theory of Viscoelasticity - An Introduction. Ed. Academic
Press, 245 pp.
Christoffersen, B., Nemat-Nasser, S., and Mehrabadi, M.M. (1981). A Micromechanical
Description of Granular Material Behavior., Journal of Applied Mechanics, 48, 339-344.
Cole, K.S., and Cole, R.H. (1941). Dispersion and Absorption in Dielectrics. I. Alternating
Current Characteristics., J.Chem.Phys., 9, 341-351.
Constable, S.C., Parker, R.L., and Constable, G.G. (1987). Occams Inversion: A Practical
Algorithm For Generating Smooth Models From Electromagnetic Sounding Data.
Geophysics, 52, 289-300.
Cundall, P.A., and Strack, O.D.L. (1979). A Discrete Numerical Model for Granular
Assemblies., Geotechnique, Vol. 29, No.1, pp. 47-65.

Bibliography

249

De Boer, R. (1996). Highlights in the Historical Development of the Porous Media Theory:
Toward a Consistent Macroscopic Theory. Applied Mechanics Review, ASME, Vol.49,
No.4, 201-262.
Delves, L.M., and Lyness, J.N. (1967). A Numerical Method for Locating the Zeros of an
Analytic Function., Math. Comp., 21, 543-560.
Dobry, R. (1970). Damping in Soils: Its Hysteretic Nature and the Linear Approximation.
Research Report R70-14, Massachusetts Institute of Technology, 82p.
Dobry, R., and Vucetic, M. (1987). Dynamic Properties and Seismic Response of Soft Clay
Deposits., Proceedings, International Symposium on Geotechnical Engineering of Soft Soils, Mexico
City, Vol.2, pp.51-87.
Drnevich, V.P. (1985). Recent Developments in Resonant Column Testing., Proceedings,
Richart Commemorative Lectures, Sponsored by Geotechnical Engineering Division, in Conjunction
with ASCE Convention, Detroit, Michigan, October 23, 1985.
Electric Power Research Institute. (1991). Proceedings: NSF/EPRI Workshop on Dynamic
Soil Properties and Site Characterization. Report NP-7337, Vol. 1, Research Project 81014.
Electric Power Research Institute. (1993). Guidelines for Determining Design Basis
Ground Motions Vol. I: Methods and Guidelines for Estimating Earthquake Ground Motion in
Eastern North America. EPRI TR-102293 Project 3302 November 1993.
Engl, H.W. (1993). Regularization Methods for the Stable Solution of Inverse Problems.
Surveys on Mathematics for Industry, Vol. 3, pp. 71-143.
Eringen, A.C. and Suhubi, E.S. (1964). Nonlinear Theory of Simple Microelastic Solid, I
and II, Int. J. Eng. Sci. 2, 189-204, 389-404.
Eringen, A.C. and Kafadar C.B. (1976). Polar Field Theories. Continuum Physics Vol. IV,
Part I, Edited by A.C. Eringen, Academic Press, p.274.
Ewing, W.M., Jardetzky, W.S., and Press, F. (1957). Elastic Waves in Layered Media.,
McGraw-Hill, 380p.
Faccioli, E., Maggio, F., Quarteroni, A., and Tagliani, A. (1996). Spectral-Domain
Decomposition Methods for the Solution of Acoustic and Elastic Wave Equations.,
Geophysics, Vol.61, No.4, pp.1160-1174.
Fam, M.A., and Santamarina, J.C. (1996). Coupled Diffusion-Fabric-Flow Phenomena: An
Effective Stress Analysis., Canadian Geotechnical Journal, Vol.33, No.3, pp.515-522.

250

Bibliography

Ferrari, M., Granik, V.T., Imam, A., and Nadeau, J.C. (1997) Advances in Doublet
Mechanics., Springer-Verlag, Berlin, 213 p.
Ferry, J.D. (1980). Viscoelastic Properties of Polymers., 3rd Edition, John Wiley, New
York, pp.641.
Finzi, B., and Pastori, M. (1961). Calcolo Tensoriale e Applicazioni, Zanichelli, Bologna,
p.508 (in Italian).
Frost, J.D., and Kuo, C.Y. (1996). Automated Determination of the Distribution of Local
Void Ratio from Digital Images., ASTM Geotechnical Testing Journal, Vol.19, No.2,
pp.107-117.
Fung, Y.C. (1965). Foundations of Solid Mechanics., Prentice-Hall, New Jersey, pp.525.
Goldstein, H. (1980). Classical Mechanics., Addison-Wesley Publishing Company, 2nd Ed.,
pp.672.
Goodman, M.A., and Cowin, S.C. (1972). A Continuum Theory for Granular Materials.
Archive for Rational Mechanics and Analysis 44: 249-266.
Granik, V.T., and Ferrari, M. (1993). Microstructural Mechanics of Granular Media.
Mechanics of Materials, 15:301-322.
Green, A.E., and Rivlin, R.S. (1964). Multipolar Continuum Mechanics., Arch. Rat. Mech.
Anal. 17, 113-147.
Gucunski, N., and Woods, R.D. (1991). Use of Rayleigh Modes in Interpretation of SASW
Tests, Proceedings, 2nd International Conference on Recent Advances in Geotechnical Earthquake
Engineering and Soil Dynamics, Vol. 2, pp. 1399-1408.
Gurtin, M.E. (1963). Variational Principles in the Linear Theory of Viscoelasticity., Arch.
Ration. Mech. Anal., 13, 179.
Hall, J.R., and Richart, Jr.F.E. (1963). Dissipation of Elastic Wave Energy in Granular
Soils., Journal of Soil Mechanics and Foundations Division, ASCE, Vol.89, No.SM6, pp.27-56.
Hardin, B.O., and Drnevich, V.P. (1972). Shear Modulus and Damping in Soils:
Measurement and Parameter Effects. J. Soil Mechanics and Foundation Engineering, ASCE,
98(SM6), 603-624.
Hardin, B.O. (1978). The Nature of Stress-Strain Behavior of Soils., Proceedings, Earthquake
Engineering and Soil Dynamics, ASCE, Pasadena, California, Vol.1, pp.3-89.
Harvey, D. (1981). Seismogram Synthesis using Normal Mode Superposition: the Locked
Mode Approximation., Geophys. J. R. Astr. Soc., 66, 37-70.

Bibliography

251

Haskell, N.A. (1953). The Dispersion of Surface Waves on Multilayered Media., Bulletin of
the Seismological Society of America, 43, 17-34.
Henrici, P. (1974). Applied and Computational Complex Analysis., Vol. 1, John Wiley &
Sons, New York, pp. 682.
Herrmann, R.B. (1994). Computer Programs in Seismology, Users Manual, Vol.II,
St.Louis University, Missouri.
Hille, E. (1973). Analytic Function Theory., Vol. 1, Chelsea, New York, 2nd Edition.
Hisada, Y. (1994). An Efficient Method for Computing Greens Functions for a Layered
Half-Space with Sources and Receivers at Close Depths, Bulletin of the Seismological Society
of America, 84(5), 1456-1472.
Hisada, Y. (1995). An Efficient Method for Computing Greens Functions for a Layered
Half-Space with Sources and Receivers at Close Depths (Part 2)., Bulletin of the
Seismological Society of America, 85(4), 1080-1093.
Holzlohner, U. (1980). Vibrations of the Elastic Half-Space Due to Vertical Surface
Loads. Earthquake Engineering and Structural Dynamics, 8, 405-414.
Idriss, I.M., and Sun, J.I. (1991). Users Manual for SHAKE91. University of California at
Davis.
Ishibashi, I. (1992). Discussion to Effect of Soil Plasticity on Cyclic Response., by M.
Vucetic and R. Dobry, Journal of Geotechnical Engineering, ASCE, Vol.118, No.5, pp.830832.
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 Behaviour in Earthquake Geotechnics., Oxford Science
Publications, Oxford, UK, pp. 350.
Iwasaki, T., Tatsuoka, F., and Takagi, Y. (1978). Shear Modulus of Sands Under Torsional
Shear Loading., Soils and Foundations, Vol.18, No.1, pp.39-56.
Jamiolkowski, M., Leroueil, S., and Lo Presti, D.C.F. (1991). Theme Lecture: Design
Parameters from Theory to Practice., Proceedings, Geo-Coast-91, Yokohama, Japan, pp.141.
Jamiolkowski, M., Lancellotta, R., and Lo Presti, D.C.F. (1994). Remarks on the Stiffness at
Small Strains of Six Italian Clays., International Symposium on Pre-FailureDeformation
Characteristics of Geomaterials, IS-Hokkaido, Sapporo, Japan, 1994.

252

Bibliography

Jenkin, C.F. (1931). The Pressure Exerted by Granular Materials: an Application of


Principle of Dilatancy. Proc. R. Soc. A. 131, 53-89.
Johnston, D.H., Toksz, M.N., and Timur, A. (1979). Attenuation of Seismic Waves In Dry
and Saturated Rocks: II. Mechanisms. Geophysics, 44(4), 691-711.
Jones, T.D. (1986). Pore Fluids and Frequency-Dependent Wave Propagation in Rocks.,
Geophysics, Vol.51, No.10, pp.1939-1953.
Jongmans, D. (1990). In-Situ Attenuation Measurements in Soils., Engineering Geology, 29,
99-118.
Kausel, E. (1981). An Explicit Solution For The Green Functions For Dynamic Loads In
Layered Media.. Massachusetts Institute of Technology, Research Report R81-13,
79
pp.
Kausel, E., and Rosset, J.M. (1981). Stiffness Matrices for Layered Soils., Bulletin of the
Seismological Society of America, 71, 6, 1743-1761.
Keilis-Borok, V.I., (1989). Seismic Surface Waves in a Laterally Inhomogeneous Earth,
Kluwer Academic Publishers, 304 pp.
Kennett, B.L.N. (1974). Reflections, Rays, and Reverberations., Bulletin of the Seismological
Society of America, 64, 1685-1696.
Kennett, B.L.N., and Kerry, N.J. (1979). Seismic Waves in a Stratified Half-Space.,
Geophys. J. R. Astr. Soc., 57, 557-583.
Kennett, B.L.N. (1983). Seismic Wave Propagation in Stratified Media., Cambridge
University Press, UK, pp.342.
Kjartansson, E. (1979). Constant Q-Wave Propagation and Attenuation., J. Geophys. Res.,
Vol.84, pp.4737-4748.
Knopoff, L. (1964). A Matrix Method for Elastic Wave Problems., Bulletin of the
Seismological Society of America, 54, 431-438.
Kokusho, T. (1980). Cyclic Triaxial Test of Dynamic Soil Properties for Wide Strain
Range., Soils and Foundations, Vol.20, No.2, pp.45-60.
Komatitsch, D., and Vilotte, J.P. (1998). The Spectral Element Method: An Efficient Tool
to Simulate the Seismic Response of 2D and 3D Geological Structures. Bulletin of the
Seismological Society of America, 88(2), 368-392.

Bibliography

253

Koppermann, S.E., Stokoe, II K.H., and Knox, D.P. (1982). Effect of State of Stress on
Velocity of Low Amplitude Compression Waves Propagating Along Principal Stress
Directions in Sand., Geotechnical Engineering Report, GR82-22, University of Texas, Austin,
Texas.
Kramer, S.L. (1996). Geotechnical Earthquake Engineering., Prentice-Hall, New Jersey,
pp.653.
Krantz, S.G. (1982). Function Theory of Several Complex Variables., John Wiley & Sons,
New York, pp.437.
Kuo, C.Y., and Frost, J.D. (1997). Initial Fabric and Uniformity of a Sand Specimen An
Image Analysis Approach., Proceedings, of ASCE Symposium on Mechanics of Deformation and
Flow of Particulate Materials, Evanston, pp. 214-227.
Kuraoka, S., and Bosscher, P.J. (1996). Parallelization of the Distinct (Discrete) Element
Method (DEM)., Symposium, Computational and Experimental Methods for Particulate
Materials, ASME, Johns Hopkins University, Baltimore, MD, June 12-14, 1996.
Lai, C.G. (1997). On the Theory of Mixtures of Porous Media., Master Report in
Engineering Science and Mechanics, The Georgia Institute of Technology, Atlanta,
Georgia.
Lamb, H. (1904). On the Propagation of Tremors over the Surface of an Elastic Solid.,
Philosophical Transactions of the Royal Society of London A203:1-42 pp.
Lambe, T.W., and Whitman, R.V. (1969). Soil Mechanics., John Wiley & Sons, New York.
Lanczos C. (1970). The Variational Principles of Mechanics., University of Toronto Press,
Toronto, 4th Edition.
Lawson, C.L., and Hanson, R.J. (1974). Solving Least Squares Problems., Prentice-Hall,
340 pp.
Lawton, W.H. and Sylvestre E.A. (1971). Elimination of Linear Parameters in Non-Linear
Regression. Technometrics, 13(3), 461-467.
Lee, W.B., and Solomon, S.C. (1979). Simultaneous Inversion of Surface-Wave Phase
Velocity and Attenuation: Rayleigh and Love Waves over Continental and Oceanic
Paths. Bulletin of the Seismological Society of America, 69(1), 65-95.
Leurer, K.C. (1997). Attenuation in Fine-Grained Marine Sediments: Extension of the BiotStoll Model by the Effective Grain Model (EGM). Geophysics, Vol.62, No.5, 1465-1479.

254

Bibliography

Liu, H.P., Anderson, D.L., and Kanamori, H. (1976). Velocity Dispersion due to
Anelasticity; Implications for Seismology and Mantle Composition., Geophys. J.R. Astr.
Soc. 47, 41-58.
Lockett, F.J. (1962). The Reflection and Refraction of Waves at an Interface between
Viscoelastic Materials., J. Mech. Phys. Solids, 10, 53.
Logan, J.D. (1997). Applied Mathematics., John Wiley & Sons, 2nd Ed., pp.476.
Lo Presti, D.C.F. (1987). Behavior of Ticino Sand During Resonant Column Tests., Ph.D.
Thesis, Politecnico di Torino, Torino, Italy.
Lo Presti, D.C.F., Jamiolkowski, M., Pallara, O. and Cavallaro, A. (1996). Rate and Creep
Effect on the Stiffness of Soils., Proceedings, Conference on Measuring and Modeling Time
Dependent Soil Behavior, Held in Conjuction with the ASCE National Convention, November 1014, 1996, Washington, D.C.
Lo Presti, D.C.F, and Pallara, O. (1997). Damping Ratio of Soils from Laboratory and InSitu Tests., Proceedings, 14th International Conference on Soil Mechanics and Foundation
Engineering, Hamburg, Germany, 6-12, September, 1997.
Lubliner, J. (1990). Plasticity Theory., Macmillan Publishing Company, New York, pp.495.
Luco, J.E. and Apsel, R.J. (1983). On the Greens Function for a Layered Half-Space. Part
I, Bulletin of the Seismological Society of America, 73, 909-929.
Lysmer, J. and Drake, L.A. (1972). A Finite Element Method for Seismology. Methods of
Computational Physics, Vol.11, Ed. Bolt, B.A., Academic press, New York, pp.181-215.
Lysmer, J. and Waas, G. (1972). Shear Waves in Plane Infinite Structures., J. Eng. Mech.
Div., ASCE, 18, 859-877.
Malagnini, L. (1996). Velocity and Attenuation Structure of Very Shallow Soils: Evidence
for Frequency-Dependent Q. Bulletin of the Seismological Society of America, 86(5), 14711486.
Malagnini, L., Herrmann, R.B., Biella, G., and De Franco, R. (1995). Rayleigh Waves in
Quaternary Alluvium from Explosive Sources: Determination of Shear-Wave Velocity
and Q Structure. Bulletin of the Seismological Society of America, 85, 900-922.
Malagnini, L., Herrmann, R.B., Mercuri, A., Opice, S., Biella, G., and De Franco, R. (1997).
Shear-Wave Velocity Structure of Sediments from the Inversion of Explosion-Induced
Rayleigh Waves: Comparison with Cross-Hole Measurements. Bulletin of the Seismological
Society of America, 87(6), 1413-1421.

Bibliography

255

Malvern, L.E. (1969). Introduction to the Mechanics of a Continuous Medium. PrenticeHall, Inc., New Jersey, 713 p.
Manolis G.D., and Beskos, D.E. (1988). Boundary Element Methods in Elastodynamics.,
Unwin Hyman, London, pp. 282.
Masad, E., Muhunthan, B., and Chameau, J.L. (1997). Stress-Strain Model for Clays with
Anisotropic Void Ratio Distribution., International Journal for Numerical and Analytical
Methods in Geomechanics, Submitted for Publication (Revised June 1997).
Menke, W. (1989). Geophysical Data Analysis: Discrete Inverse Theory., Revised Edition,
International Geophysics Series, Vol.45, Academic Press, 289 pp.
Mindlin, R.D. (1964). Microstructure in Linear Elasticity., Arch. Rat. Mech. Anal. 16, 51-78.
Mitchell, J.K. (1976). Fundamentals of Soil Behaviour., John Wiley & Sons, New York.
Muhunthan, B. (1991). Micromechanics of Steady State, Collapse and Stress-Strain
Modeling of Soils., Ph.D Thesis, Purdue University, Lafayette, Indiana, 221 pp.
Muhunthan, B., and Chameau, J.L. (1996). Void Fabric Tensor and Ultimate State Surface
of Soils., Journal of Geotechnical Engineering Division, ASCE, 123, No.2.
Nazarian, S., Stokoe, K.H., and Hudson, W.R. (1983). Use of Spectral Analysis of Surface
Waves Method for Determination of Moduli and Thicknesses of Pavement Systems.,
Transportation Research Record 930, Transportation Reseacrh Board, Washington, D.C.,
pp.38-45.
Nazarian, S. (1984). In Situ Determination of Elastic Moduli of Soil Deposits and
Pavement Systems by Spectral Analysis of Surface Waves Method., Ph.D. Dissertation,
The University of Texas at Austin.
Nemat-Nasser, S., and Mehrabadi, M.M. (1983). Stress and Fabric in Granular Mass.,
Mechanics of Granular Materials: New Models and Constitutive Relations (Jenkins, J.T., and
Satake, M. Eds.), Elsevier, pp. 1-8.
OConnell, R.J., and Budiansky, B. (1978). Measures of Dissipation in Viscoelastic Media.,
Geophys. Res. Lett., Vol.5, pp.5-8.
Oda, M. (1972). Initial Fabrics and their Relations to Mechanical Properties of Granular
Materials., Soils and Foundations, Vol.12, No.2, pp.1-18.
Oppenheim, A., and A. Willsky, A. (1997). Signals and Systems., Prentice-Hall, Englewood
Cliffs, New Jersey, 540 pp.

256

Bibliography

Papoulis, A. (1965). Probability, Random Variables, and Stochastic Processes., McGrawHill, New York, 576 pp.
Parker, R.L. (1994). Geophysical Inverse Theory., Princeton University Press, New Jersey,
pp.386.
Passman, S.L., Nunziato, J.W., and Walsh, E.K. (1984). A Theory of Multiphase Mixtures.
Appendix 5C, 286-325, Rational Thermodynamics, C. Truesdell, Springer-Verlag, Berlin,
578 pp.
Pipkin, A.C. (1986). Lectures on Viscoelasticity Theory., 2nd Edition, Springer-Verlag,
Berlin, pp.188.
Press, W.H., Teukolsky, S.A., Vetterling, W.T., and Flannery, B.P. (1992). Numerical
Recipies in Fortran - The Art of Scientific Computing., Cambridge University Press, 2nd
Ed., pp. 963.
Read, W.T. (1950). Stress Analysis for Compressible Viscoelastic Materials., J. Appl. Phys.,
21, 671.
Richart, F.E., Jr., Woods, R.D., and Hall, J.R. (1970), Vibrations of Soils and Foundations.,
Prentice-Hall, Englewood Cliffs, New Jersey, 414 pp.
Rix, G.J. (1988). Experimental Study of Factors Affecting the Spectral Analysis of Surface
Waves Method., Ph.D. Dissertation, The University of Texas at Austin, pp.315.
Rix, G.J., Lai, C.G., Spang, A.W.,Jr. (1998a). In-Situ Measurement of Damping Ratio Using
Surface Waves. Accepted for publication to ASCE Journal of Geotechnical and
Geoenvironmental Engineering, 29 pp.
Rix, G.J., and Lai, C.G. (1998). Simultaneous Inversion of Surface Wave Velocity and
Attenuation, Geotechnical Site Characterization, Edited by P.K. Robertson and P.W.
Mayne, Vol. 1, pp.503-508, Proceedings of the First International Conference on Site
Characterization ISC98/Atlanta, Georgia, USA, 19-22 April 1998.
Rix, G.J., Lai, C.G., Foti, S., and Zywicki D. (1998b). Surface Wave Tests in Landfills and
Embankments, Proceedings, 3rd ASCE Conference on Soil Dynamics and Earthquake Engineering
and Soil Dynamics Conference, Seattle, Washington, USA, August, 3-6,1998.
Rosset, J.M., Chang, D.W., Stokoe, K.H. II (1991). Comparison of 2-D and 3-D Models
for Analysis of Surface Wave Tests. 5th International Conference on Soil Dynamics and
Earthquake Engineering, Karlsruhe, Germany, 1991, pp. 111-126.
Rothenburg, L. (1980). Micromechanics of Idealized Granular Systems., Ph.D Thesis,
Carleton University, Ottawa, Canada, 332 pp.

Bibliography

257

Snchez-Salinero, I. (1987). Analytical Investigation of Seismic Methods Used for


Engineering Applications., Ph.D. Dissertation, The University of Texas at Austin, pp.
401.
Satake, M. (1982). Fabric Tensor in Granular Materials., IUTAM Symposium On Deformation
and Failure of Granular Materials, Delft, pp. 63-68.
Schwab, F., and Knopoff, L. (1970). Surface-Wave Dispersion Computations., Bulletin of
the Seismological Society of America, 60, 321-344.
Schwab, F., and Knopoff, L. (1971). Surface Waves on Multilayered Anelastic Media.,
Bulletin of the Seismological Society of America, 61, 4, 893-912.
Schwab, F., and Knopoff, L. (1972). Fast Surface Wave and Free Mode Computations.,
Methods of Computational Physics, Vol.11, Ed. Bolt, B.A., Academic press, New York,
pp.87-180.
Seed, H.B., and Idriss, I.M. (1970). Soil Moduli and Damping Factors for Dynamic
Response Analyses., Report EERC 70-10, Earthquake Engineering Research Center,
University of California, Berkeley.
Shibuya, S., Mitachi, T., Fukuda, F., and Degoshi, T. (1995). Strain Rate Effects on Shear
Modulus and Damping of Normally Consolidated Clay. Geotechnical Testing Journal, 18(3),
365-375.
Spang, A.W., Jr. (1995). In Situ Measurements of Damping Ratio Using Surface Waves.
Ph.D. Dissertation, Georgia Institute of Technology, pp. 347.
Stoll, R.D. (1974). Acoustic Waves in Saturated Sediments. Physics of Sound in Marine
Sediments, Plenum Press, 19-39.
Stokoe, K.H. II, Rix, G.J., and Nazarian, S. (1989). In Situ Seismic Testing with Surface
Waves., Proceedings, 12th International Conference on Soil Mechanics and Foundation Engineering,
Rio De Janeiro, 13-18 August, pp. 331-334.
Takeuchi, H. and Saito, M. (1972). Seismic Surface Waves. Methods of Computational Physics,
Vol.11, Ed. Bolt, B.A., Academic press, New York, pp.217-294.
Tang, X.M.. (1992). A Waveform Inversion Technique for Measuring Elastic Wave
Attenuation Using Cylindrical Bars. Geophysics, 57, 854-859.
Thomson, W.T. (1950). Transmission of Elastic Waves through a Stratified Solid Medium,
J.Appl.Phys., 21, 89-93.
Tikhonov, A.N. and Arsenin, V.Y. (1977). Solutions of Ill-Posed Problems., Winston &
Sons, Washington D.C., pp.258.

258

Bibliography

Ting, J.M., Corkum, B.T., Kauffman, C.R., and Green, C. (1989). Discrete Numerical
Model for Soil Mechanics., J. Geotechnical Engineering, ASCE, 115 (3), 379-398.
Ting, J.M., Khwaja, M., Meachum, L.R., and Rowell, J.D. (1993). An Ellipse-Based Discrete
Element Model for Granular Materials., International Journal for Numerical and Analytical
Methods in Geomechanics, V. 17, pp. 603-623.
Tokimatsu, K. et al. (1992). Effects of Multiple Modes on Rayleigh Wave Dispersion
Characteristics. J. Geotechnical Engineering, ASCE, 118(10), 1529-1543.
Tokimatsu, K. (1995). Geotechnical Site Characterization using Surface Waves. Proceedings,
First International Conference on Earthquake Geotechnical Engineering, IS-Tokyo '95, Tokyo,
November 14-16, Balkema, Rotterdam, 1333-1368.
Truesdell, C. (1957). Sulle Basi della Termomeccanica., Accademia Nazionale dei Lincei,
Rendiconti della Classe di Scienze Fisiche, Matematiche e Naturali (8), 22, 33-88, 158-166 (in
Italian).
Truesdell, C. and Noll W. (1992). The Non-Linear Field Theories of Mechanics., SpringerVerlag, 2nd Edition, p.591.
Tschoegl, N.W. (1989). The Phenomenological Theory of Linear Viscoelastic Behavior An Introduction., Springer-Verlag, Berlin, pp.769.
Visintin, A. (1994). Differential Models of Hysteresis., Springer-Verlag, Berlin, pp.407.
Vrettos, C. (1991). Time-Harmonic Boussinesq Problem For a Continuously NonHomogeneous Soil., Earthquake Engineering and Structural Dynamics, 20, 961-977.
Vucetic, M., and Dobry, R. (1991). Effect of Soil Plasticity on Cyclic Response. J.
Geotechnical Engineering, ASCE, 117(1), 89-107.
Vucetic, M. (1994). Cyclic Threshold Shear Strains in Soils., Journal of Geotechnical
Engineering, ASCE, Vol.120, No.12, pp.2208-2228.
Wilmanski, K. (1996). The Thermodynamical Model of Compressible Porous Materials
with the Balance Equation of Porosity, Journal of Non-Equilibrium Thermodynamics, 21, pp.
1-30.
Winkler, K.W., and Nur, A. (1979). Pore Fluids and Seismic Attenuation in Rocks.,
Geophysical Research Letters, Vol. 6, 1-4.
Zeng, Y., and Anderson, J.G. (1995). A Method for Direct Computation of the Differential
Seismogram with Respect to the Velocity Change in a Layered Elastic Solid., Bulletin of
the Seismological Society of America, Vol.85, No.1, pp.300-307.

Das könnte Ihnen auch gefallen