You are on page 1of 213

A HYBRID EFFECTIVE STRESS TOTAL STRESS

PROCEDURE FOR ANALYZING SOIL EMBANKMENTS


SUBJECTED TO POTENTIAL LIQUEFACTION AND FLOW

by
Ernest Naesgaard

B.A.Sc., The University of British Columbia, 1973


M.Eng., The University of British Columbia, 1988

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF


THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in

THE FACULTY OF GRADUATE STUDIES

(Civil Engineering)

THE UNIVERSITY OF BRITISH COLUMBIA


(Vancouver)

April 2011

Ernest Naesgaard, 2011


Abstract
Seismic design of major civil structures (bridges, dams and embankments) is
moving increasingly towards using performance design methodologies which require
determination of earthquake induced movements. Development of these numerical design
tools and procedures for use in engineering practice for estimating the earthquake induced
ground deformations of potentially liquefiable soil is the topic of this dissertation.

Fully coupled effective stress numerical analyses procedures developed at the


University of British Columbia (UBC) were used to simulate field and centrifuge test case
histories. These analyses can offer considerable insight, but due to the complexity of the
problem and variability of the parameters involved, there is considerable uncertainty. The
author, therefore, recommends that the relatively new state-of-the-art effective stress
analyses should be augmented by carrying out an additional analysis compatible with
conventional design processes. This latter analysis uses published post-liquefaction
residual soil strengths derived from back-analysis of field case histories by others. The
developed design methodology uses the effective stress (UBCSAND) soil constitutive model
for dynamic analyses, and empirical residual post-liquefaction soil strengths for a post-
shaking total stress static analysis. In the proposed approach, the effective stress dynamic
analysis is used to determine zones of liquefaction, to quantify earthquake induced
deformations, and to provide overall insight. The post-shaking total stress static analysis,
with residual strength parameters used in elements which liquefied, is carried out to
capture the effects of complex stratigraphy and localization that may be missed by the
effective stress model.

Calibration and validation of the UBCSAND model was undertaken by comparing


the model with field case histories and laboratory simple shear, shake table, and centrifuge
tests. The measured response of some centrifuge tests being used for validation was
indicative of the centrifuge model not being fully saturated. This was problematic as P-wave
measurements within the centrifuge model suggested full saturation. A series of triaxial tests
with P-wave measurements was carried out. These tests, and the numerical modeling of
them, showed that high P-wave velocities were not always indicative of full saturation and
they provided a logical explanation for the observed centrifuge response.

ii
Preface
Technical advance evolves over time and involves the contribution of many. This
research is no different.

Chapter 2 is an overview of my understanding of key aspects of soil liquefaction


and related to phenomena. Material included in chapter was previously published in
Naesgaard et al. (2005, 2006, 2007, and 2009) with co-authors including Professor P.M.
Byrne, Dr. M. Seid-Karbasi, and Dr. M. Beaty. I wrote the papers with input and review
from the co-authors. Much insight also comes from the work of Professor T. Kokusho.

The material in Chapter 3, Soil Mixing, was previously published by Naesgaard


and Byrne (2005). Professor Byrne proposed the mixing concept in 1988/89 to explain the
flow of tailings into the underground Mufulira mine workings. I proposed the concepts of
cyclic mixing as a means of inducing soil liquefaction and low strength and flow within
sands. Others, Vasquez-Herrera and Dobry (1989), had similar thoughts. I carried out
rudimentary cyclic shear tests to study the mixing of layered granular soils. I also made the
observation that mixing could be induced by turbulent flow in the vicinity of water films.

Chapter 4 is a brief overview on post-liquefaction residual strength that is based on


literature review.

Chapter 5 is largely a discussion on the constitutive models UBCSAND, UBCTOT,


and UBCHYST that were developed at the University of British Columbia by Professor
Byrne and his students. UBCTOT was developed by Professor Byrne and Dr. Beaty.
UBCSAND was developed by Professor Byrne and several students. I made modifications
to UBCSAND described in Section 5.3.4 and compiled a list of limitations and challenges
with the program. The algorithm for UBCHYST was developed by Professor Byrne. I
coded UBCHYST as a constitutive model in FLAC, made improvements regarding
unload/reload loops, developed options for strength reduction as a function of strain or
number of cycles, and developed documentation explaining the model.

The proposed hybrid effective stress total stress design approach in Chapter 6 is a
new contribution derived from my research and is published in papers by Naesgaard, Byrne,
and Beaty (2009), and by Naesgaard and Byrne (2007).

iii
Chapter 7 contains UBCSAND model calibrations and example analyses that I have
either carried out or supervised, unless otherwise indicated.

Section 7.2 provides a summary of recent single element numerical simulations of the
simple shear tests by Sriskandakumar (2004).
Section 7.3 describes a calibration of the model to empirical liquefaction triggering
charts. This procedure is commonly used for calibrating UBCSAND for commercial
design projects. The example shown was calculated by Dr. Beaty.
Section 7.4 describes some early 2004 one dimensional (infinite slope) column tests that
I carried out to demonstrate that UBCSAND/FLAC could simulate the post-shaking
flow failure phenomena. This may be the first simulation of this in a numerical model.
Section 7.5 describes one of the earlier commercial uses of UBCSAND with a centrifuge
test program for validation. The work is described in Naesgaard et al. (2004) and in
Yang et al. (2004) with co-authors P.M. Byrne, K. Adalier, T. Abdoun, and B. Gohl.
Section 7.6 is from a paper by Naesgaard, Byrne, Seid-Karbasi, and Park (2005) that
described some numerical modeling of centrifuge tests carried out as part of the UBC
Liquefaction Initiative. I carried out the numerical modeling described in this section.
Section 7.7 describes numerical modeling that I conducted to simulate the behaviour
observed in Professor Kokushos shake table tests. The work was previously published
in a paper by Naesgaard, Byrne, and Beaty (2009).
Section 7.8 describes my work on numerical modeling of the Lower San Fernando dam
in California. This was previously published in Naesgaard and Byrne (2007). This
section compares an actual well documented field case history to that obtained using the
proposed design methodology. Other, practicing engineers, have also adopted the
models described in this dissertation and are using them on similar projects to not only
provide insight into liquefaction phenomena, but also to quantify earthquake induced
ground movements.
The work described in Chapter 8 was previously published by Naesgaard, Byrne
and Wijewickreme (2007). I did the laboratory testing and numerical analyses, while
Professor Byrne and Professor Wijewickreme provided consultation and review.

E. Naesgaard
April 2011

iv
Table of Contents
ABSTRACT .......................................................................................................................ii
PREFACE ........................................................................................................................ iii
TABLE OF CONTENTS .................................................................................................. v
LIST OF TABLES ......................................................................................................... viii
LIST OF FIGURES .......................................................................................................... ix
LIST OF SYMBOLS .....................................................................................................xvii
ACKNOWLEDGEMENTS ............................................................................................ xx
DEDICATION ................................................................................................................ xxi
1 INTRODUCTION........................................................................................................ 1
2 SOIL LIQUEFACTION, VOID REDISTRIBUTION AND POST-
LIQUEFACTION RESIDUAL SHEAR STRENGTH ............................................ 3
2.1 Overview ............................................................................................................. 3
2.2 Loose Sand Layer underlying Low Permeability Barrier ................................... 7
2.3 Pore Water Void Redistribution within real Embankments and Slopes ........... 10
2.4 Chapter 2 Summary and Conclusions ............................................................... 14
3 SOIL MIXING ........................................................................................................... 25
3.1 Introduction to Mixing Phenomena .................................................................. 25
3.2 Mufulira Mine Case History (mixing due to static slope failure) ..................... 26
3.3 Cyclic Mixing in Rudimentary Laboratory Shear Tests ................................... 27
3.4 Discussion ......................................................................................................... 28
3.5 Chapter 3 Summary and Conclusions ............................................................... 30
4 RESIDUAL STRENGTH FROM BACK-ANALYSES OF CASE
HISTORIES................................................................................................................ 43
4.1 Introduction ....................................................................................................... 43
4.2 Seed (1984, 1987) ............................................................................................. 45
4.3 Seed and Harder (1990) .................................................................................... 45
4.4 Olson and Stark (2002) ..................................................................................... 45
4.5 Idriss and Boulanger (2008) .............................................................................. 45
4.6 Discussion ......................................................................................................... 46
4.7 Chapter 4 Summary and Conclusions ............................................................... 46

v
5 ANALYSIS PROCEDURES OVERVIEW ............................................................. 51
5.1 Introduction ....................................................................................................... 51
5.2 State-of-Practice Total Stress Analysis (UBCTOT) ......................................... 53
5.3 UBCSAND Coupled Effective Stress Analysis ................................................ 54
5.3.1 Introduction to UBCSAND ............................................................................... 54
5.3.2 Elastic Response................................................................................................ 56
5.3.3 Plastic Response ................................................................................................ 57
5.3.4 Recent Improvements/Options added to UBCSAND ....................................... 58
5.3.4.1 Pull Down Yield Surface on unloading if no Cross-over .................... 58
5.3.4.2 Modeling of Dense Soils with Lower Phase Transformation
Friction Angle...................................................................................... 58
5.3.4.3 Delay Dilation following Post-Liquefaction Cross-over ..................... 58
5.3.4.4 Dilation Reduction or Cut-off upon Expansion of Element ................ 59
5.3.5 UBCSAND/FLAC Limitations and Challenges ............................................... 60
5.3.6 UBCSAND/FLAC Grid-Model Development Considerations ......................... 63
5.4 Hysteretic Model for Non-liquefiable Clay/Silt Soils (UBCHYST) ................ 64
5.4.1 Introduction ....................................................................................................... 64
5.4.2 Non-Linear Hysteretic UBCHYST Model .................................................... 65
5.4.3 Implementation of UBCHYST Model .............................................................. 66
5.4.4 UBCHYST Calibration ..................................................................................... 67
5.4.5 Summary ........................................................................................................... 67
5.5 Chapter 5 Summary and Conclusions ............................................................... 67
6 PROPOSED COMBINED EFFECTIVE STRESS TOTAL STRESS
APPROACH ............................................................................................................... 79
6.1 Introduction ....................................................................................................... 79
6.2 Implementation of Procedure in FLAC............................................................. 79
6.3 Discussion ......................................................................................................... 84
7 MODEL CALIBRATION, VALIDATION AND APPLICATION
EXAMPLE.................................................................................................................. 89
7.1 Introduction ....................................................................................................... 89
7.2 Calibration to Laboratory Cyclic Simple Shear Tests....................................... 89
7.2.1 Introduction ....................................................................................................... 89
7.2.2 Results and Discussion...................................................................................... 90
7.3 Calibration to Seed type Empirical Liquefaction Triggering Charts ............. 91

vi
7.4 One-Dimensional Tests for demonstrating Void Redistribution ...................... 93
7.5 George Massey Tunnel Centrifuge Test Modeling ........................................... 94
7.6 UBC/C-CORE Centrifuge Test Modeling ........................................................ 95
7.7 Kokusho Shake Table Emulation ...................................................................... 96
7.8 Lower San Fernando Dam Example Analysis .................................................. 97
7.8.1 Introduction ....................................................................................................... 97
7.8.2 Coupled Effective Stress Analysis of LSFD Phases 2 and 3a ....................... 98
7.8.3 Post-Liquefaction Total Stress Analysis of LSFD Phase 3b .......................... 98
7.8.4 Results of LSFD Example Analysis .................................................................. 99
8 P-WAVE MEASUREMENTS AS AN INDICATOR OF SATURATION
IN CENTRIFUGE TESTS ...................................................................................... 131
8.1 Introduction ..................................................................................................... 131
8.2 Laboratory Element Testing Program ............................................................. 133
8.3 Numerical Modeling ....................................................................................... 135
8.4 Discussion ....................................................................................................... 136
8.5 Chapter 8 Summary and Conclusions ............................................................. 137
9 DISCUSSION AND CONCLUSIONS ................................................................... 149
9.1 Summary and Conclusions .............................................................................. 149
9.2 Recommendations ........................................................................................... 156
9.2.1 Recommendations based on the Work carried out for this Dissertation ......... 156
9.2.2 Recommendations for Future Research .......................................................... 157
REFERENCES .............................................................................................................. 160
APPENDIX A UBCSAND1v02 FISH code and Flow Chart ....................................... 169
APPENDIX B Analysis Of A Concrete Gravity Dam Over Potential Liquifiable
Soil Illustrating Proposed Hybrid Procedure Methodology .................. 184

vii
List of Tables
Table 3-1 Summary of Rudimentary Cyclic Shear Tests................................................ 33
Table 7-1 Summary of Undrained Laboratory Simple Shear and UBCSAND
Simulations .................................................................................................... 101
Table 7-2 Massey Tunnel Project. Class A Numerical Predictions compared
to Centrifuge Test Results (in prototype scale) ............................................. 101
Table 7-3 Soil properties used in emulation of Kokusho shake table tests . ...................102
Table 8-1 Summary of Laboratory Test Conditions and Results................................... 140
Table 8-2 Soil and Fluid Properties used in Numerical Analyses ................................. 141
Table B-1 Example Dam analysis model profile and soils types. Soil
properties are summarized in Table B-1 ........................................................ 189

viii
List of Figures
Figure 2-1 Comparison of cyclic drained simple shear response of loose and
dense Fraser River sand (from Sriskandakumar 2004). (a) is shear
stress versus shear strain, (b) is shear strain versus volumetric strain,
and (c) is shear stress versus volumetric strain. (Note labels in (c)
refer to the loading phase the soil is always contractive on
unloading). .......................................................................................................17
Figure 2-2 Cyclic (a) stress path and (b) stress-strain response of loose sand
without initial static shear stress (from Sriskandakumar 2004). ......................18
Figure 2-3 Postulated response of loose sand to undrained cyclic simple shear
loading followed by monotonic loading to large strain. (a)
is stress strain space, (b) is stress path space and (c) is void ratio
vertical effective confining stress space. A is initial state with pore
pressure ratio, R u =0, D is true liquefaction with R u = 1, and F is
the strength at the critical or steady state if pore water does not
cavitate. ............................................................................................................19
Figure 2-4 Postulated response of loose sand element to cyclic simple shear with
A to B undrained and B to D with slight pore water inflow.
(a) is stress strain space, (b) is stress path space and (c) is void ratio
vertical effective confining stress space. A is initial state with pore
pressure ratio, R u =0, D is true liquefaction with R u = 1. At D the
sample is at the critical state with strength at or near
zero........................................................................................................ .......... 20
Figure 2-5 Re-consolidation volumetric strain from dissipation of excess pore
pressure following liquefaction. = 0 is with zero static bias or level
ground conditions whereas = static /' vo = 0.1 is with a static bias
such as that from sloped ground (from triaxial data by Dismuke 2003
as reported by Malvick 2005). .........................................................................21
Figure 2-6 Hammer blow to base of 130mm diameter column of loose sand
liquefies it and causes almost immediate formation of water film at
boundary with less permeable fine sand (Kokusho and Kabasawa
2004). ...............................................................................................................21
Figure 2-7 Path A is the stress path for a soil element with a static bias and with
pore water inflow that is cyclically loaded until flow instability
occurs. Path C would be an element with there is pore water outflow
and flow instability does not occur (Malvick 2005 after Kulasingam
2003). ...............................................................................................................22
Figure 2-8 Factors affecting thickness and residual strength of localized shear
zone under barrier (a) grain-size (b) leakage through barrier (c)
undulations of barrier surface (d) variation of stress on barrier (from
Naesgaard et al. 2006). .....................................................................................23

ix
Figure 2-9 Flow redistribution mechanisms within an embankment with low
permeability barrier and high permeable layer. Vectors show
directions of pore water flow (Naesgaard et al. 2006). ....................................24
Figure 3-1 Graphic illustrating the effects of soil mixing. Elements 'A', 'B' and
'C' have the same number and volume of large and small particles.
A is the element with a fine layer over a coarse layer prior to
disturbance. 'B' represents the case where fine particles fall into the
voids between large particles and leave a water filled void or
interlayer at the top of the element. C represents the case where the
particles mix evenly and the larger particles move into a looser
arrangement. Both B and C will have significantly lower shear
strength then A if mixing does not occur (Naesgaard and Byrne
2005). ...............................................................................................................35
Figure 3-2 1970s Mufulira Mine tailings liquefaction failure. .........................................35
Figure 3-3 Photograph of 1970 Mufulira Mine sink hole (tailings.info 2005). .................36
Figure 3-4 Grain-size of Mufulira Mine tailings (Barrett and Byrne 1988;
Horsfield and Been 1989). ...............................................................................36
Figure 3-5 Void ratio vs. undrained residual shear strength ( res = (1-3)/2)
relationship for sandy silt, silty sand, and mixed residual in-rush
tailings. Note: Mixed tailings have essentially zero shear strength at
the in-situ void ratio (Naesgaard and Byrne 2005 from triaxial tests
by Horsfield and Been 1989). ..........................................................................37
Figure 3-6 Conceptual undrained triaxial compression stress path for silty sand,
sandy silt and mixed silt-sand at same void ratio and initial stress state
(Naesgaard and Byrne 2005). ...........................................................................37
Figure 3-7 Postulated progressive failure mechanism of 1970 sinkhole at
Mufulira Mine. (1) Sides of initially step sided sinkhole fails, (2) as
soil slides to bottom of slope, the layers mix and due to this mixing
lose nearly all strength (liquefy), (3) the liquefied soil flows away into
mine workings, (4) mechanism is repeated giving a progressive
failure mechanism, and (5) the final stable slope at shallow 10 angle
(Naesgaard and Byrne 2005). .......................................................................... 38
Figure 3-8 Rudimentary cyclic shear test with saturated layered sample
(Naesgaard and Byrne 2005). ...........................................................................38
Figure 3-9 Grain-size of fine and coarse sand used in rudimentary simple shear
tests (Naesgaard and Byrne 2005)....................................................................39
Figure 3-10 Test A before and after cyclic shearing (Naesgaard and Byrne
2005). ...............................................................................................................39
Figure 3-11 Post-shearing for non-layered Tests B and C (Naesgaard and
Byrne 2005). .....................................................................................................40

x
Figure 3-12 Before and after cyclic shearing for Test D with glass beads and
coarse sand (Naesgaard and Byrne 2005). .......................................................40
Figure 3-13 Slide from Kokusho cylinder shake table test video showing zone of
mixing between coarse and fine sand caused by pore fluid flow and
turbulence (Kokusho 2003) (Video from: http//:www.civil.chuo-
u.ac.jp/lab/doshitu/index.html).........................................................................41
Figure 3-14 Steady state void ratio vs. residual strength (Sus) for reconstituted
(remolded) Lower San Fernando Dam (California) soil. Note that the
residual strength (S us ) for the combined (mixed) homogeneous
specimens is much lower than that of the layered not-mixed
specimens (Baziar and Dobry 1995). (Without mixing the strength is
about 0.31 tsf whereas with mixing the strength is about 0.06 tsf). .................42
Figure 4-1 Post-liquefaction residual strength form case histories (a) Seed et. al.
(1984), (b) Seed and Harder (1990), (c) Olsen and Stark (2002) and
(d) Idriss and Boulanger (2008). ......................................................................49
Figure 4-2 Post-liquefaction residual strength form case histories by Idriss and
Boulanger (2008) with undrained strength inferred from University of
British Columbia laboratory tests by Sriskandakumar (2004)
superimposed (Byrne et al., 2008). ..................................................................50
Figure 5-1 In the equivalent linear method initial G(1) and (1) are estimated.
An elastic analysis is carried out with these values and from the
results a new (1) eff is obtained. This (1) eff is used to get a new G(2)
and (2) which are used for the second elastic analysis iteration. This
process is repeated until the process converges and strain compatible
GFinal and final are obtained. ......................................................................... 70
Figure 5-2 Example of the liquefaction triggering methodology in UBCTOT.
The CSR at A would cause liquefaction in 15 cycles; however,
since it is only a cycle this pulse provides 1/15 x = 1/30th of the
loading to cause liquefaction. Pulse B would cause liquefaction in
2 cycles and, therefore, provides x = of the loading necessary
to cause liquefaction. The full cycle of A and B would provide
+ 1/30 = 17/60 of the loading necessary to liquefy the element. ...................... 70
Figure 5-3 Bilinear response of an element that liquefies when using UBCTOT
(Beaty, 2001). ...................................................................................................71
Figure 5-4 Liquefaction triggering chart and equations from Idriss and Boulanger
(2006, 2008). ....................................................................................................72
Figure 5-5 Yield surface in UBCSAND. The hardener (d = d(/')) that
expands the yield surface (i.e., from A to B) is a function of the
plastic shear strain and plastic shear modulus as illustrated in Figure
5-8. ...................................................................................................................72
Figure 5-6 Flow Rule in UBCSAND. Below cv or pt shear strain induces
volumetric contraction and above it induces volumetric dilation. ...................73

xi
Figure 5-7 Stress ratio history showing definition of loading, unloading and
reloading UBCSAND (Beaty 2009; Beaty and Byrne 2011)...........................73
Figure 5-8 Plastic strain increment and plastic modulus in UBCSAND. ..........................74
Figure 5-9 Comparison of postulated post-liquefaction response of loose sand
and that modeled by UBCSAND. ....................................................................74
Figure 5-10 A dilation delay (proportional to dilation (expansive plastic
volumetric strain) in last cycle) is induced at shear stress cross-over
in UBCSAND...................................................................................................75
Figure 5-11 Large element(c) emulating the behaviour of a small element (b). If
(a) and (b) are at critical state with l/l = L/L then (c) can be made to
behave similarly to (b) by setting dilation to 0 when the volumetric
strain is equal to l/L ........................................................................................75
Figure 5-12 Loading in simple shear induces hysteretic response in soil with shear
modulus (G) and loop-size (damping) varying with strain. .............................76
Figure 5-13 Characteristic behaviour of soil in cyclic shearing. .........................................76
Figure 5-14 UBCHYST model key variables. .....................................................................77
Figure 5-15 Example calibration of UBCHYST constitutive model to G/G max and
damping ratio curves. .......................................................................................77
Figure 5-16 Comparison in response between laboratory simple shear test and
UBCHYST simulation for case with static bias (note that the scales
on plots are not the same).................................................................................78
Figure 6-1 Flow chart showing combined coupled effective stress / total stress
analysis procedure (Naesgaard and Byrne 2007). ............................................88
Figure 7-1 Typical grain-size distribution and microscope view of Fraser River
Sand used for cyclic shear tests (from Sriskandakumar 2004). .....................104
Figure 7-2 Comparison of undrained simple shear test L1 laboratory test to
UBCSAND simulation (a) pore pressure ratio vs. number of cycles,
(b) stress path, and (c) stress strain response. ................................................105
Figure 7-3 Comparison of undrained simple shear test L2 laboratory test to
UBCSAND simulation (a) pore pressure ratio vs. number of cycles,
(b) stress path, and (c) stress strain response. ................................................106
Figure 7-4 Comparison of undrained simple shear test L2 laboratory test to
UBCSAND simulation (a) pore pressure ratio vs. number of cycles,
(b) stress path, and (c) stress strain response. ................................................107
Figure 7-5 Comparison of Undrained simple shear test L12 laboratory test with
initial static bias to UBCSAND simulation (a) pore pressure ratio vs.
number of cycles, (b) stress path, and (c) stress strain response
using same calibration parameters as those used without static bias. ............108

xii
Figure 7-6 Comparison of undrained simple shear test L12 laboratory test with
initial static bias to UBCSAND simulation (a) pore pressure ratio vs.
number of cycles, (b) stress path, and (c) stress strain response
using additional hfac7 calibration parameter to control pull-down
prior to triggering. ..........................................................................................109
Figure 7-7 Comparison of cycles to liquefaction between undrained laboratory
simple shear tests and UBCSAND simulations. ............................................110
Figure 7-8 Undrained dense sand simple shear stress vs. strain comparison of D3
laboratory test to UBCSAND simulation (a) Laboratory test, and (b)
UBCSAND simulation. ..................................................................................111
Figure 7-9 Undrained dense sand simple shear stress path comparison of D3
laboratory test (a) to UBCSAND simulation (b)............................................112
Figure 7-10 Undrained dense sand simple shear pore pressure ratio comparison of
D3 laboratory test (a) to UBCSAND simulation (b). .....................................113
Figure 7-11 Empirical liquefaction triggering chart with overlain UBCSAND
simulations (from Beaty 2009; Beaty and Byrne 2011).................................114
Figure 7-12 Volumetric strain (v) within an infinite slope column with a low
permeability crust over loose sand. ................................................................115
Figure 7-13 UBCSAND numerical model of infinite slope (1D) column with low
permeability barrier cap. At time x-x' zone 'I' has expanded to the
critical state, dilation goes to zero, and flow failure is initiated. ...................115
Figure 7-14 Section through Massey Tunnel in Fraser River. ...........................................116
Figure 7-15 Typical transverse FLAC model grid used from Massey Tunnel.
Scale: the tunnel is 24m wide - vertical and horizontal scale are the
same................................................................................................................116
Figure 7-16 Typical distorted mesh with displacement vectors. Scale: the tunnel is
24m wide - vertical and horizontal scale are the same...................................116
Figure 7-17 Typical distorted mesh with displacement vectors for level ground
condition. Scale: the tunnel is 24m wide - vertical and horizontal
scale are the same. ..........................................................................................117
Figure 7-18 Layout of Massey project centrifuge model showing location of
instrumentation Model 2 (Legend: a3 = accelerometer; P12 =
piezometer; Lv and Lh= LVDT). Dimensions in centimeters. Model
1 and Model 3 are similar...............................................................................117
Figure 7-19 Model #1 accelerometer A10 (dark line is FLAC class A prediction
and light line is from centrifuge test). ............................................................118
Figure 7-20 Comparison of class A numerical prediction and centrifuge results for
Model #2 with 10m wide densification. .........................................................118

xiii
Figure 7-21 Comparison of pore pressure in piezometer P5 in Model #3 (dark
colour is FLAC class A prediction and lighter colour is centrifuge test
result)..............................................................................................................119
Figure 7-22 Typical grid and input time history used for UBC Liquefaction
Initiative C-CORE centrifuge tests (shown in prototype scale). ....................119
Figure 7-23 Comparison of Centrifuge tests and numerical results for profile with
low permeability silt barrier (COSTA-C): (a) displaced grid, (b)
Horizontal displacement of sliding block over barrier (for the
centrifuge data the solid line is measured and the dashed line is
corrected to better match final displacements), (c) vertical
displacement at crest, (d) centrifuge and numerical surface profiles,
(e) calculated lateral displacement contours in metres, (f) and (g) pore
pressure time histories at P3 and P6, (h) acceleration time history at
A6. ..................................................................................................................120
Figure 7-24 (a) Initial and displaced profile of centrifuge test CT5 with three
drainage slots (b) numerical analysis of same................................................121
Figure 7-25 Comparison of vertical displacement near crest with (CT5) and
without (COSTA-C) drainage slots. Post-shaking flow initiated in the
COSTA-C test at approximately 70s..............................................................122
Figure 7-26 Kokusho shake table model at start of test (Kokusho 2003). .........................122
Figure 7-27 Kokusho shake table test profile at end-of-shaking. ......................................123
Figure 7-28 Post-failure profile of Kokusho shake table test. Failure occurred a
short time after all shaking had stopped. ........................................................123
Figure 7-29 UBCSAND/FLAC model grid of Kokusho shake table test. .........................124
Figure 7-30 Original boundary of model and displaced shape of FLAC model
grid. ................................................................................................................124
Figure 7-31 Contours of horizontal displacement (a) with barrier layers in place
and flow on and (b) with barriers but no flow. ..............................................125
Figure 7-32 Calculated displacement time histories of shake table model with and
without barrier and with and without flow. ....................................................126
Figure 7-33 Lower San Fernando Dam which failed into the reservoir
approximately 30s after end of earthquake shaking. Blue areas are
soil that was mixed and deemed to have liquefied. Lower figure is
reconstructed profile which shows location of liquefied soil before
failure. (Seed et al. 1973). ..............................................................................127
Figure 7-34 Grid, locations of low permeability barriers, (N1) 60 and shear strength
input parameters used in Lower San Fernando numerical example. .............128
Figure 7-35 Profile showing zones deemed to have liquefied at end of strong
shaking (R u > 0.7). .........................................................................................129

xiv
Figure 7-36 Displacement time histories of point near upstream toe of dam for
both 3a coupled effective stress analysis (ESA) and 3b total stress
analysis (TSA) alternative. Both analyses stopped due to excessive
element distortion. ........................................................................................129
Figure 7-37 (a) Horizontal velocity (m/s) and (b) horizontal displacement (m) at
end of analyses (40s) for coupled effective stress analysis. .........................130
Figure 8-1 Variation of (a) Skempton B with Degree of Saturation (S r ) and (b)
P-wave velocity (V p ) with Skempton B for Fraser River Sand, with
D r =20% K b =92 MPa, G=56 MPa, and porosity = 0.45 in accordance
with Equations (1) and (2). ...........................................................................143
Figure 8-2 Normalized cyclic strength ratio (CSR) of sand compared to P-wave
velocity (V p ) as measured in laboratory tests conducted with
Toyoura sand. Developed using data from Tsukamoto et al. (2002)
and Ishihara et al. (1998). CSR is that which causes liquefaction
(double amplitude strain of 5%) in 20 cycles (for Relative Density of
40-70%). .......................................................................................................143
Figure 8-3 Laboratory test layout. ..................................................................................144
Figure 8-4 Variation of shear wave velocity (V s ) with effective mean normal
confining pressure ( m ) as measured in laboratory tests. ............................144
Figure 8-5 Typical P-wave time histories as measured in the laboratory. Traces
(a) is for B = 0.48 with a calculated V p of 1415 m/s and trace (b) is
for B = 0.95 with a calculated V p of 1557 m/s. .............................................145
Figure 8-6 Measured shear wave (V s ) and P-wave (V p ) versus Skempton B. .............146
Figure 8-7 35 by 75 element mesh used for FLAC analyses. ........................................146
Figure 8-8 (a) Represents homogeneous-partial-saturation (HPS) where the air
bubbles are small and scattered throughout the void spaces. This is
numerically modeled by reducing the fluid modulus (K w ) at all the
nodes in the model. (b) Represents non-homogeneous-partial-
saturation (NHPS) where there are only a few large bubbles at select
locations. This is numerically modeled by reducing the fluid
modulus only at select nodes within the model. ...........................................147
Figure 8-9 Comparing V p versus Skempton B from both HPS and NHPS FLAC
analyses to that observed in laboratory tests and that calculated from
Equation (2). Data from Ishihara et al. (1998) and Tamura et al.
(2002) that deviated from the correlation from Equation (2) are also
shown. ............................................................................................................ 147
Figure 8-10 Typical variation of fluid modulus within NHPS FLAC grid. Small
squares represent locations with low fluid modulus representative of
air bubbles, whereas all other elements have a high fluid modulus of
2.2x109 Pa. ....................................................................................................148

xv
Figure A-1 Flow diagram for UBCSAND1v02 illustrating how it fits within the
FLAC (ITASCA 2008) Mohr Coulomb framework. ......................................18
Figure B-1 Example Dam analysis model profile and soils types. Soil properties
are summarized in Table B-1. .......................................................................189
Figure B-2 Example Dam analyses model grid details. ..................................................190
Figure B-3 Example Dam analysis typical post-earthquake horizontal
displacement contours. .................................................................................190
Figure B-4 Example Dam analyses showing typical post-earthquake vertical
displacement contours. .................................................................................191
Figure B-5 Example Dam analysis showing typical post-earthquake pore
pressure ratio contours. .................................................................................191
Figure B-6 Time histories showing displacement of center of intake structure
and model base (top of till) for typical analysis (positive is
downstream or upward). ...............................................................................192

xvi
List of Symbols
B: Skemptons coefficient for pore pressure
Be : Elastic bulk modulus
CRR: cyclic resistance ratio
CRR 15 : cyclic resistance ratio
CSR: cyclic stress ratio
Dr: relative density
D 50 : mean size of soil grains (50% by weight passing)
e: void ratio
e max : maximum void ratio
e min: minimum void ratio
G: shear modulus
G max : Shear modulus at small strain
Ge: elastic shear modulus
GP: plastic shear modulus
G liq : liquefied soil shear modulus
Gp i : initial plastic modulus at =0
i: hydraulic gradient
K o : at-rest soil pressure coefficient
K: bulk modulus
K: elastic bulk modulus
K w : bulk modulus of fluid (water)
K mix : bulk modulus of air-fluid mixture
Ke G : shear modulus number
Ke B : bulk modulus number
K : correction factor for static bias or slope
K : correction for confining pressure
K b: bulk modulus of soil skeleton
K m : earthquake magnitude correction factor
M: constraint bulk modulus
m_hfac1 to m_hfac7: UBCSAND calibration factors
Mod1: UBCHYST calibration parameter,
- a reduction factor for first-time or virgin loading
Mod2:UBCHYST calibration parameter
- to account for permanent modulus reduction with large strain (optional)
Mod3: UBCHYST calibration parameter
- to account for cyclic degradation of modulus with strain or number of cycles, etc.
n: porosity
n: UBCHYST calibration parameter
n 1: UBCHYST calibration parameter
(N 1 ) 60 : normalized Standard Penetration Test, N-value for energy and overburden
(N 1 ) 60-cs : normalized N-value for clean sand
(N 1 ) 60cs-Sr : SPT N value normalized and fines corrected for use in calculating residual shear
strength
OCR: over-consolidation ratio
xvii
Pa: atmospheric pressure
PI: soil plastic limit
PGA: peak horizontal ground acceleration
Q c1cs : normalized and fines corrected CPT tip bearing pressure
Rf: failure stress ratio
R u : ratio of excess pore pressure to initial vertical effective stress
R f : UBCHYST calibration parameter constant that truncates hyperbolic curve
Sr: saturation
S r : post-liquefaction residual strength
Sss: steady state strength
SPT: Standard Penetration Test
t: time
TL: thickness of liquefied soil layer beneath barrier layer
u: pore pressure
ue: excess pore pressure
V p : compression wave velocity
V s : shear wave velocity
t: time
$switch: a 0/1 flag indicating whether or not the post liquefaction volumetric strain has
reached the desired v
u: change in pore pressure
V : decrement of vertical stress
: static bias
p s : plastic shear strains
: strain
1 : axial strain
a : axial strain
e: elastic strain
p: plastic strain
p v : Plastic volumetric
v : desired volumetric strain in the liquefied element according to published correlation
v : volumetric strain
: stress ratio
1 : change in stress ratio
1f : change in stress ratio to reach failure envelope in direction of loading
1fold : 1f in previous reversals
1old : 1 in previous reversals
f : stress ratio at failure
max : maximum stress ratio () at last reversal
: Poisons ratio
= density
': effective stress
' m : mean normal effective stress
tot : total stress
V: vertical stress
' v : vertical effective stress

xviii
v : vertical total stress
'vo: effective initial vertical stress
: shear stress
: shear stress
res: residual strength
xy : developed shear stress in horizontal plane
xy : shear stress on the horizontal ( xy ) plane
: shear strain
: unit density
e: elastic shear strain
max : maximum shear strain during earthquake shaking
p: plastic shear strain
w: water unit weight
: soil friction angle
cs/ss: quasi/steady state friction angle
cv : constant volume friction angle
f : Peak friction angle
f : peak friction angle
p : peak friction angle
pt : phase transformation friction angle
: dilation angle

xix
Acknowledgements

The financial support from the British Columbia Ministry of Transportation - UBC
Professional Partnership Program, the Canadian Council of Professional Engineers, Trow
Associates Inc., and the National Scientific and Engineering Research Council Strategic
Liquefaction Grant No. NSERC 246394 is gratefully acknowledged. The triaxial testing
system with bender elements were purchased using a grant from The Canada Foundation for
Innovation (CFI) - New Opportunities Infrastructure Funding, Project No. 6869. Assistance
from BC Hydro is also much appreciated.

Discussion and support from colleagues and friends Ali Amini, Michael Beaty,
Jusheng Qian, Ali Khalili, Mavi Sanin, Pascale Rouse, Mahmood Seid Karbasi, David Siu,
Alex Sy, Sung Sik Park, (Uthaya) M. Uthayakumar, Somasundaram Sriskandakumar, Sonny
Singha, James Wetherill and many others is much appreciated, as is the drafting assistance
from Danny Lam, and formatting and editing input from Diane McCulloch and Julie Sedger.
Correspondence with Peter Cundall of ITASCA and Ryan Phillips of C-CORE on soil
saturation and P-wave correlations and modeling is also greatly appreciated.

The author would like to thank the University of British Columbia staff and faculty
for their assistance and enlightenment. The author wishes to thank his research committee
members Dr. Upul Atukorala and Professor Mahdi Taiebat, and most importantly his
supervisors, Professor Peter M. Byrne and Professor Dharma Wijewickreme, for their
support and encouragement.

xx
DEDICATION

to Tineke, Daniel, Jasper, Mom, my late Father


and
Everyone else who made this happen

xxi
Chapter 1 - Introduction

1 INTRODUCTION
Seismic design of major civil structures (bridges, dams and embankments) is moving
increasingly toward the use of performance design methodologies. The owner, design code,
or other body holding jurisdiction generally specifies performance criteria for the proposed
structure. Then it is the designers task to demonstrate that the structure will perform in
accordance with the design criteria. A key aspect of performance design is to determine the
earthquake induced ground and structure movements. To do this, designers are increasingly
resorting to the use of dynamic numerical analyses. When the soils or rock involved are
dense, these analyses can often be carried out using relatively simple elastic-plastic
constitutive models. However, with saturated loose granular soils there is potential for soil
liquefaction, related large deformations and possible flow sliding. This behaviour is
significantly more complex, and the numerical procedures required reflect this. The research
for this dissertation involves development of these numerical design tools and procedures for
use in engineering practice for estimating the earthquake induced ground deformations of
soil that may potentially liquefy. The developed design methodology uses effective stress
and total stress soil constitutive models, UBCSAND and UBCHYST, developed at the
University of British Columbia (UBC). There are several assumptions and approximations
in these analysis procedures and due to this, empirical soil strengths that have been back-
calculated from past earthquake case histories are also used as part of the process.

Chapters 2 and 3 of this dissertation give an overview of the behaviour of saturated


loose sandy soils, prone to liquefaction during strong earthquake shaking. Chapter 2
discusses liquefaction and pore water void redistribution mechanisms and Chapter 3 focuses
on soil-mixing as an alternative means for low strengths and flow. Chapter 4 discusses
empirical post-liquefaction soil strengths derived from field case histories. Chapter 5
presents an overview of the numerical analyses models while Chapter 6 discusses suggested
analysis methodology for use in design work. Related model calibration and analysis
examples are given in Chapter 7. Chapter 8 provides a diversion from the main dissertation
theme. Some of the work described in Chapter 7 involved centrifuge testing at C-CORE,
cyclic simple shear laboratory testing at UBC (Sriskandakumar 2004), and numerical
modeling at UBC and Memorial University of Newfoundland (University of British

1
Chapter 1 - Introduction

Columbia Liquefaction Initiative research program (NSERC Grant 246394)). Both


centrifuge tests and laboratory tests were used for numerical model calibration purposes, and
are described in Chapter 7. The response of some of the centrifuge tests was indicative of the
centrifuge model not having been fully saturated. This was puzzling as P-wave
measurements within the centrifuge model were indicative of full saturation. A series of
triaxial tests with P-wave measurement apparatus were carried out at UBC. These tests
showed that high P-wave velocities were not always indicative of full saturation (Naesgaard
et al. 2007) and provided an explanation of the observed centrifuge response. A summary of
this testing program and subsequent conclusions is presented in Chapter 8. An overview,
conclusions and recommendations for further research are presented in Chapter 9.

2
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

2 SOIL LIQUEFACTION, VOID REDISTRIBUTION and POST-


LIQUEFACTION RESIDUAL SHEAR STRENGTH
2.1 Overview

An overview of the characteristic behaviour of sands when sheared monotonically,


cyclically, drained, and undrained has been thoroughly described by Seid Karbasi (2008) and
numerous other authors (Idriss and Boulanger 2008; Byrne et al. 2006; Vaid and Chern 1985;
Castro 1975 and Casagrande 1936). This will not be reiterated in this dissertation and only
an overview of the key aspects related to the proposed modeling is presented. This overview
is the authors interpretation and understanding of key points.

Fundamentally, the response of soil is related to the response of particle to particle


contacts to internal pore fluid pressures and external boundary stresses. Mathematical and
numerical models, such as, PFC3D (ITASCA 2003), that simulate the particulate behaviour
are available. These models currently have limited ability but are developing quickly and
with advancement in computing power are expected to become important design tools in the
near future. Current methodologies, including those discussed in this dissertation, generally
treat the soil as a continuum. The intent is that the constitutive model of a single element in
the continuum should simulate the soil skeleton pore fluid interaction with external stresses
that are observed in laboratory element tests. When these elements are combined into a
group, with suitable compatibility and continuity equations, will simulate the behaviour
observed in large physical models and field scenarios. In this chapter, both the behaviour of
a single element (drained, undrained, and partially drained response as observed in a simple
shear laboratory test) and the behaviour of a soil profile, or in numerical terms a group of
elements, are discussed. The purpose of these discussions is to provide a background on
behaviour inherent in potentially liquefiable soil deposits. This behaviour should, ideally, be
emulated by the proposed numerical model, and understood by the users of the numerical
model.

Drained Response: When typical loose sand, relative density (D r = 40%), is tested in
drained (dry) cyclic simple shear, its stress path is similar to that experienced by a one
dimensional soil column during earthquake loading. Initially, the soil is contractive on
loading and unloading. However, on loading to large strains, the stress state exceeds a phase

3
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

transformation stress ratio (approximately equal to the tangent of the constant volume
friction angle ( cv ) in loose sands) and the soil becomes dilative (Figure 2-1) (Ishihara 1993;
Vaid et al. 1981). Dense sand (D r = 80%) behaves in a similar manner except the dilative
response is more pronounced and phase transformation occurs at lower shear strain (Figure
2-1) and sometimes at lower stress ratio. Both dense and loose sands are contractive on
unloading. In most cases, the net result of drained cyclic loading in both dense and loose
sands is a reduction in sample volume (Finn et al. 1982).

Undrained Response: If the soil pores are filled with water that has been prevented from
escaping (undrained condition), then pore pressures will increase when the soil skeleton
attempts to contract, and decrease when the soil skeleton attempts to expand (dilate). With
repeated cycles the net result is generally an increase in pore pressure and related decrease in
effective stress (Vaid and Chern 1985; Sriskandakumar 2004). With a sufficient number of
cycles, the stress path may reach the zero effective stress (' = 0) and zero shear strength
( = 0) origin in the - ' stress space, and true liquefaction occurs (Figure 2-2). However,
with continued shearing following true liquefaction, the soil will usually attempt to dilate and
gain strength (Figure 2-2) (Vaid and Chern 1985). At very large strains, the soil will reach a
critical or steady state where further shear strain does not induce pore pressure or volume
change and further strength gain will not occur (Casagrande 1936; Negussey et al. 1988).
The terms critical state and steady state are the same state and are taken to be synonyms in
this dissertation. It is postulated that in undrained loading, the post-liquefaction strength of a
soil element will not be reached until, (i) the pore water cavitates and, thus, allows the
sample to increase in volume (dilate) and reach the critical state, or (ii) the high mean
effective stress generated by dilation suppresses the dilation and the soil reaches its critical
state strength, or (iii) the sand grains crush and the soil reaches a critical state of the crushed
material (Figure 2-3). The strength of the sand reached in (i), (ii) or (iii) is generally much
larger than the commonly accepted post-liquefaction undrained strengths back-calculated
from case histories (Seed 1987; Seed and Harder 1990; Olson and Stark 2002; and Idriss and
Boulanger 2008) and is often larger than the drained strength. In this dissertation, these
back-calculated post-liquefaction strengths are referred to as the residual strength ( res or
S r ) of the liquefied soil.

4
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

Response with small inflow (partially drained condition): If, in lieu of undrained loading,
a small inflow of water (expansion) occurs, then shear can occur without the strength gain
observed in undrained tests. Upon reaching the critical state, further inflow will rapidly
cause total loss of strength (Vaid and Eliadorani 1998; Boulanger and Truman 1996; and
Sento et al. 2004) (Figure 2-4). This process of void or pore water redistribution was
proposed by Whitman (1985) (NCR 1985) and received recent attention within the
engineering design community by the shake table tests by Kokusho (1999, 2003). As
illustrated in Figure 2-4, with continued water inflow, the shear strength of a soil element
will always go to zero (water film condition).

Residual Strength in the Field: Natural soils and many man-made soils are often layered
and have variations in grain-size and permeability. Earthquake shaking and related
liquefaction will induce pore water gradients and flow that will result in contraction of the
soil skeleton with outflow of pore water in some areas, and expansion with inflow of pore
water, in other areas (void redistribution) (NCR 1985; Whitman 1985; Naesgaard et al.
2006). Inflow produces soil strengths considerably lower than those obtained assuming
undrained conditions (Vaid and Eliadorani, 1998). When there is a low permeability layer or
barrier, the migrating pore water becomes trapped beneath the barrier and forms an area of
local expansion with low effective stress (NCR 1985). With sufficient inflow, an actual
water interlayer with zero shear strength will develop. The residual strength back-
calculated from field case histories commonly used for design is not a fundamental soil
property nor the strength of a single soil element, but rather the average strength of multiple
elements that are deemed to have liquefied. This is discussed further in Section 2.3.

Consolidation and other volumetric strain mechanisms within liquefied soil: Volume
reduction of the soil skeleton of saturated granular soils subjected to earthquake shaking or
other cyclic loading is postulated to occur by several distinct mechanisms.

Mechanism 1 is net volume reduction and/or pore water pressure increase


induced by cyclic shear strain and related particle rearrangement (Taylor
1948, Rowe 1962);

5
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

Mechanism 2 is volume reduction of the soil skeleton due to sedimentation of


particles and only occurs at those times in the process when the mean effective
stress is essentially zero (Kokusho et al. 2003; Malvick 2005; Dashti 2009);

Mechanism 3 is plastic consolidation of the soil skeleton with mean effective


stress greater than zero (this often occurs in combination with Mechanism 4)
(Terzaghi and Peck 1967);

Mechanism 4 is elastic consolidation of the soil skeleton due to soil loading


(increase in mean effective stress) (Terzaghi and Peck 1967); and,

Mechanism 5 is reduction of the skeleton volume due to mixing of layers of


coarse and fine soil (likely more prevalent at low effective stresses) (Byrne
1989, Naesgaard and Byrne 2005).

Typical net volume reduction resulting from earthquake shaking and post-
earthquake consolidation in areas of relatively flat topography (small initial static bias) are in
the range of 2-5% for looser sands, and in the range of 0.5% for denser sands (Malvick 2002;
Idriss and Boulanger 2008). Observations from physical testing indicates that during cyclic
loading much of the particle movement and settlement occurs during the portion of the cycles
when the soil is passing through the shear stress/effective stress origin (zero effective stress)
and is postulated to take place mainly by Mechanism 2 (Kokusho et al. 2003). When there
are significant shear stresses present, the sedimentation (Mechanism 2) does not take place
and volume reduction is typically significantly reduced (Malvick 2005). This is illustrated in
triaxial test data by Dismuke (2003) (Malvick 2005) and shown in Figure 2-5.

Stress Redistribution: When granular soils within an embankment are subjected to


earthquake shaking, the shear may induce contraction of the soil skeleton in some areas and
corresponding build-up of pore pressure and reduction in effective stress. This reduction in
effective stress will not occur uniformly throughout the embankment and will be more
prevalent in some areas than others. When the effective stress decreases within a local zone
or pocket, the shear modulus of the soil in that zone or pocket will also decrease, and shear
stresses that were previously supported by this soil will be shed to adjacent stiffer zones
(Naesgaard et al. 2006). Thus, even though the overall static bias within an embankment
slope may be significant, there will still be local zones or pockets of liquefied soil with

6
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

limited, or no static bias. With reduced shear stresses, the soil in these pockets will be more
contractive, effective stresses lower, and the ability of that zone to resist shear stresses will
be further reduced.

2.2 Loose Sand Layer underlying Low Permeability Barrier

In this Section, the postulated behaviour of a one-dimensional column of loose (D r = 40%)


sand underlying an impermeable barrier layer is discussed.

It is assumed that the initial shaking is sufficient to produce true liquefaction (zero
effective stress) throughout the soil underlying the barrier for the level ground Case A. For
Cases B and C with a static bias, it is assumed that the shaking is sufficient to reduce the
effective stress to the minimum value required to support the shear stress associated with the
static bias. Also, for Cases B and C, it is assumed that the static bias is lower than the phase
transformation friction angle so that shaking produces a net increase in pore pressure. When
referring to shear stress () in this Section it is the shear stress on a plane parallel to the
ground surface and not necessarily the maximum shear stress ( max ). Effective stress is the
mean normal effective stress (' m ) unless indicated otherwise.

Case A Level ground (no static bias): At the on-set of earthquake shaking, the full depth
of the loose sand layer attempts to contract, pore pressures increase and effective stresses
decrease. During the shaking cycles, there may be alternating shear induced contraction and
dilation, and also some expansion of the sand skeleton due to a reduction in effective stress.
However, the net result is contraction of the skeleton and an increase in pore pressure
throughout the column. In a short time, a pore pressure gradient develops with the highest
head at the base of the column and a lower head at the top. This causes pore water to flow
upward relative to the soil particles. A soil element at the base of the column has a net
outflow of pore water and consolidates relatively quickly, whereas, soil elements between the
top and bottom of the column experience little volume change, as water flowing into and out
of each element is similar. At the top of the sand column, directly under the impermeable
barrier, water flows out. It cannot pass through the impermeable barrier, and a water
interlayer develops. If the barrier is totally impermeable, the water incompressible, and the
sand layer is totally liquefied (i.e., effective stress is zero), then a water interlayer develops
almost immediately. With time, outflow from the bottom soil element diminishes, and the

7
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

next higher element (above the bottom element) consolidates because the outflow now
exceeds the inflow. This process advances as a consolidation front, moving from the bottom
upward. During the consolidation process, the size of the water layer under the barrier
increases. It should be noted that the shear strength of a water interlayer is zero.

In addition to settlement due to sedimentation, there will be plastic and elastic


settlement due to the restoration of effective stress. The restoration of effective stress does
two things, (1) it causes greater consolidation at the bottom layers due to the change in
effective stress being greater with depth, and (2) it causes the flow gradient to decrease more
in the bottom parts of the layer than in the top parts. The net result of (1) and (2) is greater
contraction at the base of the layer and the potential for some expansion in the upper soil
layers that are below the barrier water film as proposed by Seid-Karbasi (2008). For very
loose sands, it is postulated that sedimentation-type consolidation dominates and that
settlement due to the restoration of effective stress is small as the bulk modulus of sands
increases rapidly with effective stress. For example, assuming an effective stress change
from 0 to 100 kPa and a constrained modulus of 500Pa(' m / Pa)0.5 (where Pa = 100 kPa)
then by integration a volumetric strain due to elastic settlement will be = 0.004 or 0.4%.
Typical post-liquefaction settlement from both field case histories and laboratory tests is
about 2% to 5% (Idriss and Boulanger 2008) and is about 10 times higher than the elastic
portion.

In summary, the net result of earthquake shaking on loose sand, which underlies an
impermeable barrier, in level ground conditions (no static bias), will be contraction
throughout the height of the deposit. A water interlayer will form at the top of the sand
column. The thickness of this water interlayer will be equal to the net consolidation in the
underlying loose sand and would, typically, range from 2 to 5%, depending on density. This
is illustrated by the post-impact formation of localized water films in columnar tube tests by
Kokusho (Kokusho and Kabasawa 2003) (Figure 2-6). The water film illustrated in Figure
2-6 occurred almost immediately after the base of the column was struck. Note that if the
overlying fine sand in Figure 2-6 was removed, the remaining coarse sand would still settle
in a similar manner when struck, but the water would pond on the surface, rather than form
an interlayer.

8
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

The inherent challenges of modeling the above process numerically are discussed in
Chapter 5. For denser sands, the skeleton volume reduction due to sedimentation would be
lower and settlement due to restoration of effective stress would be a larger portion of the
total settlement.

Case B Sloping ground with stress path during shaking that pass through the - ' m
origin: At the on-set of earthquake shaking, the full depth of the loose sand soil column
attempts to contract, pore pressures increase, and effective stresses decrease.

(i) Assuming the sand is dense of the critical state (a common situation even for loose
sands (Vaid and Thomas 1995)), then the effective stress does not reduce to zero
except during cross-over when the stress path passes through the - ' m origin.
(ii) Sedimentation settlement (Mechanism 2) only occurs during shaking and only
during cross-over (when ' m is zero) the amount of settlement that occurs during
shaking is largely a function of permeability. The sand will have essentially zero
bulk and shear modulus during the cross-over and a large cyclic mobility and
movement in direction of static bias may occur during shaking.
(iii) Post-shaking vertical or mean effective stress will be limited by the value required
to support the static bias and will not go to zero (unless there is sufficient pore water
inflow to take the sand to the critical state (see (v) below) (Boulanger and Truman
1996).
(iv) Much of the post-shaking settlement will be due to reinstatement of effective stress
(Mechanisms 3 and 4) and will be significantly less than that which would occur in
the absence of static bias (as indicated in Figure 2-5 by Path C1-C3 with static bias
compared to path W0-W3 without bias).
(v) With time, the pore pressure gradient will reduce with depth, hence, inflow may
exceed outflow in the upper sections. The extreme upper boundary of the sand unit
has no outflow due to the impermeable barrier, and, therefore, expands more than
the other zones. Expansion reduces shear strength (both because the effective
stresses drop, and because the friction angle is decreasing from a peak value toward
a critical state value), and shear modulus which leads to further localization. With
sufficient inflow, the upper element reaches the critical state and further inflow

9
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

causes flow failure. Failure will occur just before a water film forms due to the
static bias (Malvick 2005).

Case C Sloping ground with stress path during shaking not passing through the -
' m origin: At the on-set of earthquake shaking, the full depth of the loose sand soil column
attempts to contract, pore pressures increase and effective stresses decrease; however,
effective stresses do not go below that required to support the static bias (Boulanger and
Truman 1996, Malvick 2005, Kulasingam 2003).

(i) Assuming the sand is dense of the critical state (a normal condition), then initially
the effective stress does not reach zero and is sufficient to support the shear stress in
the soil.
(ii) Decrease in effective stress leads to some cyclic mobility, especially the case if the
stress path intersects the failure plane, however, the deformations will often be less
than that for Case B, where the cycles pass through the stress path origin.
(iii) A pore pressure gradient will still form within the column. Lower parts of the
column will experience a net loss of water (as indicated by Path C in Figure 2-7)
and maintain or regain strength whereas soil in the upper part of the column under
the barrier will gain the water expelled from the lower parts and expand (as
indicated by Path A in Figure 2-7). As indicated by Path A in Figure 2-7, once
inflow is sufficient to take the sample to the critical state then shear localization
occurs and further inflow leads to flow failure.
(iv) Sedimentation settlement (Mechanism 2) will be essentially zero and post-shaking
settlements will be due to reinstatement of effective stress only. Post-shaking
settlements will be significantly less than those which would occur for Case A with
no static bias and less than that which would occur for Case B where there are cross-
overs.

2.3 Pore Water Void Redistribution within real Embankments and Slopes

In real soils, the barrier boundary will not be perfectly plane, of infinite lateral extent, nor
will the localization be infinitely thin, nor the barrier totally impermeable. In a real world
scenario, there will be undulations, varying normal stress, finite grain-sizes, etc. (Figure 2-8)
(Naesgaard et al. 2005, 2006). These items will result in the residual shear strength along

10
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

the interface varying both in time and space and will usually result in an average value that is
greater than zero. Some key factors postulated to affect the shear strength and behaviour of
the barrier interface include:

(i) Soil density - Dense soil (at a given confining pressure) has greater affinity for
expansion and can accommodate more inflow before reaching the critical state, than
does a loose soil (at the same confining pressure) (Seid-Karbasi 2008).

(ii) Net volume of inflow - The available pore water inflow will be a function of the
volume and relative density of the loose layer feeding the inflow, permeability of
the loose layer, pore water flow path, and, earthquake shaking amplitude and
duration. Dense soil will contract less and, therefore, have less pore water to
contribute to the expansion beneath the barrier than a loose soil.

(iii) Grain-size - The larger the grains the greater the thickness of the shear band and the
amount of inflow required for steady state shear to be achieved in the shear band
(Figure 2-8(a)). Roscoe (1970) suggests that localized shear bands are in the order
of 10 to 20 times the particle mean diameter (D 50 ).

(iv) Permeability and continuity of the barrier - Leakage through the barrier will reduce
the net inflow into the interface (Figure 2-8(b)).

(v) Waviness or undulations of barrier interface - These may be undulations that


precede earthquake shaking and/or may be due to deformations induced by strong
earthquake shaking during the event (Figure 2-8(c)).

(vi) Variations in vertical stress - Variations in vertical total stress along the barrier
boundary will allow the higher stressed sections (Figure 2-8(d) X) to maintain
some strength while the lower stressed sections (Figure2-8(d) Y) heave.

(vii) Shear stress redistribution will occur and instead of there being an area with uniform
static shear stress bias, stronger points will attract stress, and other areas will have
low stress with possibly no resulting static bias. The areas with no static bias may
have essentially zero shear strength due to water film development and possibly
have significant sedimentation settlement, whereas, zones attracting shear will have
finite residual shear strengths and significantly less consolidation settlement.

11
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

The above conditions will result in the residual shear strength along the interface
varying in both time and space, resulting in an average value that is greater than zero. This
average value is the residual shear strength that is believed to be reflected in the low
values back-calculated from case histories.

When inflow is into a zone with a static shear bias a reduction in effective stress will
initially occur with nominal shear strain until the zone hits the failure envelope (in stress path
space). Once on the failure envelope a large shear strain that is a function of inflow volume
and dilation angle will occur (Boulanger and Truman 1996, Sento et al. 2004) while the shear
stress is maintained at the value of the static bias. With continued inflow, the layer will
eventually reach the critical state at a shear stress corresponding to the static bias, and any
further inflow will lead to strength loss and flow slide. This is illustrated in Figure 2-7.

In the one dimensional infinite slope model previously discussed in Section 2.2, the
net pore water migration is essentially vertical. However, in embankments and water-edge
slopes both vertical and/or lateral flow may be responsible for pore water void redistribution
effects (NRC 1985). Figure 2-9 illustrates pore water void redistribution concepts within a
hypothetical embankment of loose sand with a single low permeability barrier layer and a
high permeability layer. Only flow failure in the upstream direction is being considered.
Initially, liquefaction will tend to be more prevalent and trigger earlier near the centre of the
embankment where the static bias is lower. The on-set of liquefaction (shaded area around
A and B in Figure 2-9(a)) initiates a complex series of reactions. The author postulates
these to include:

(1) The liquefied zone will soften and shed load toward the edges of the embankment,
thus reducing static bias within the liquefied zone and increasing the shear stress on
the outer zones.

(2) The deeper portions of the liquefied soil (A in Figure 2-9(a)) will have a higher
hydraulic head than that of the overlying soil (B in Figure 2-9(a)). Both A and
B will have a higher hydraulic head than that within the shells of the embankment
(C in Figure 2-8(a)). This will initiate upward and outward pore water flow. The
outer zones of the embankment are often dilative will also attempt to draw water in
from the reservoir.

12
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

(3) The upward rising pore water will become trapped under the low permeability
barrier at B in Fig 2-9, cause the soil in the vicinity of the barrier to expand, and
form a localized weak layer and possible water film at the underside of the barrier.

(4) The expansion of the soil under the barrier will cause an increase in permeability at
the underside of the barrier which will facilitate a progressive lateral migration of
water into the highly stressed outer zones of the embankment.

(5) Pore water migrating into the outer embankment zones (both from the central
portions of the dam and the reservoir) will reduce effective confining stresses and
potentially bring the soil in the shell to failure. Once on the failure envelope a shear
strain approximately proportional to the inflow induced volumetric expansion will
ensue. This will continue until the inflow subsides or the localized layers expand to
the critical state (Sento et al. 2004). Once at the critical state, further inflow causes
loss of all strength and possibly flow failure.

(6) Lenses of high permeability will also facilitate the migration of pore water from the
higher head central portions of the embankment toward the lower head and highly
stressed shells. The effects of this are as discussed in (5).

(7) The on-set of liquefaction generally occurs during strong earthquake shaking.
However, pore water void redistribution takes time and will occur both during and
after earthquake shaking. The process may lead to the triggering of a flow slide
(during or after end-of-shaking) or cause limited movements that stabilize due to
geometry changes or because the source of water driving the expansion is no longer
present.

(8) Higher volumetric strains and shear localization will occur immediately below
barriers where there is pore water inflow. With localization the inflow (volume of
water) required for the soils to go to the critical state and to initiate flow failure is
reduced.

(9) In embankments and water-edge slopes, liquefaction within the centre portion of
embankments or back from a water-edge slope will induce high pore pressures
approaching the local overburden stress. This initial liquefaction results in (1) the

13
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

liquefied zones shedding shear stress to adjacent non-liquefied zones, and (2) the
generation of local high pore pressures and pore pressure gradients. The
redistribution of shear stress reduces the static bias within the liquefied soils
allowing effective stresses to go essentially to zero and allowing greater potential
for volume change. The high pressure water migrates both vertically and laterally
towards the edges of the embankment, or towards the water-edge slope (location of
lower overburden pressure), whereby expansion and potential for strength reduction
will also occur. Since water flow and pressure redistribution takes time, there is
often a delay between occurrence of earthquake shaking and time of flow failure.

There are numerous case histories where soil liquefaction occurred during
earthquake shaking, but related flow failure did not occur until sometime after the shaking
terminated. Documented examples include the Lower San Fernando Dam, California, where
the upstream face of the dam failed 20 to 30 seconds after the end of earthquake shaking
(Seed 1973; Seed 1987), and Niigata, Japan, where eyewitnesses reported that the girders of
the Showa Ohashi Bridge fell a few minutes after the earthquake motion had ceased
(Hamada 1992).

2.4 Chapter 2 Summary and Conclusions

In this Chapter, an overview of items contributing to liquefaction and potential flow sliding
are given. These are behaviours that are important for the proposed numerical model to
capture and important for the proposed users of the model to understand.

It is noted that sands have contractive tendencies when sheared at stress ratio less
than the phase transformation stress ratio, have dilative tendencies sheared at higher stress
ratio, but always contractive tendencies on unloading. The net response of cyclic shear
loading of saturated sands is nearly always a tendency for contraction. This in turn generates
pore water pressures, potential liquefaction of the sands (low effective stress and very low
shear moduli), and sets up a hydraulic gradient within the deposit. The gradients cause water
flow (generally upward or toward a free face) out of some zones and into others. Those
zones losing water generally contract and those gaining pore water expand. When there is a
low permeability barrier above the sand layer the inflow of pore water into the soil directly
below the barrier results in localization of both the expansion and shear strain. As the sand

14
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

expands, its potential shear-induced-dilative-tendencies decrease, and with sufficient


expansion, the critical state, at which shear strain will induce no further expansion, is
reached. This contraction and expansion can have a dramatic affect on both the effective
stress and the shear strength of the sand. This is particularly the case for loose sand under
low permeability barriers where expansion tends to be localized and strengths can go to very
low values, or at the limit, zero when a water interlayer forms.

Part of the behaviour described in this Chapter concerns that of a soil element (i.e.,
drained and undrained response) and the other part is related to the interaction between zones
or groups of elements in numerical terms (i.e., water flow, void and stress redistribution).
This distinction is noted when modelling the behaviour numerically using coupled effective
stress methods. The soil constitutive model captures the effective stress-strain element
behaviour of the soil skeleton, while a suitable computational process captures the coupling
of volume changes within the skeleton with pore water pressure, pore water flow, and stress
strain interaction of the whole model (Byrne et al. 2006; Taiebat et. al. 2010). Modelling of
the above behaviour using the constitutive model UBCSAND (Beaty and Byrne 1998; Byrne
et al. 2004) within the commercially available program FLAC (ITASCA 2008) is further
discussed in Chapters 5 to 7.

Most of the concepts in this Chapter have been discussed by others at some time in
the past. However, the framework and emphasis of the discussion is important and is that of
the author.

15
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

FIGURES

Chapter 2
Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual
Shear Strength

16
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

Figure 2-1 Comparison of cyclic drained simple shear response of loose and dense Fraser River
sand (from Sriskandakumar 2004). (a) is shear stress versus shear strain, (b) is shear
strain versus volumetric strain, and (c) is shear stress versus volumetric strain. (Note
labels in (c) refer to the loading phase the soil is always contractive on unloading).

17
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

Figure 2-2 Cyclic (a) stress path and (b) stress-strain response of loose sand without initial
static shear stress (from Sriskandakumar 2004).

18
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

Figure 2-3 Postulated response of loose sand to undrained cyclic simple shear loading
followed by monotonic loading to large strain. (a) is stress strain space, (b) is
stress path space and (c) is void ratio vertical effective confining stress space.
A is initial state with pore pressure ratio, Ru=0, D is true liquefaction with Ru
= 1, and F is the strength at the critical or steady state if pore water does not
cavitate.

19
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

Figure 2-4 Postulated response of loose sand element to cyclic simple shear with A to B
undrained and B to D with slight pore water inflow. (a) is stress strain space,
(b) is stress path space and (c) is void ratio vertical effective confining stress space.
A is initial state with pore pressure ratio, Ru=0, D is true liquefaction with Ru =
1. At D the sample is at the critical state with strength at or near zero.

20
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

Figure 2-5 Re-consolidation volumetric strain from dissipation of excess pore pressure
following liquefaction. = 0 is with zero static bias or level ground conditions
whereas = static/'vo = 0.1 is with a static bias such as that from sloped ground
(from triaxial data by Dismuke 2003 as reported by Malvick 2005).

Figure 2-6 Hammer blow to base of 130mm diameter column of loose sand liquefies it and
causes almost immediate formation of water film at boundary with less permeable
fine sand (Kokusho and Kabasawa 2004).

21
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

Figure 2-7 Path A is the stress path for a soil element with a static bias and with pore water inflow
that is cyclically loaded until flow instability occurs. Path C would be an element with
there is pore water outflow and flow instability does not occur (Malvick 2005 after
Kulasingam 2003).

22
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

Figure 2-8 Factors affecting thickness and residual strength of localized shear zone under
barrier (a) grain-size (b) leakage through barrier (c) undulations of barrier surface
(d) variation of stress on barrier (from Naesgaard et al. 2006).

23
Chapter 2 Soil Liquefaction, Void Redistribution and Post-Liquefaction Residual Shear Strength

Figure 2-9 Flow redistribution mechanisms within an embankment with low permeability barrier
and high permeable layer. Vectors show directions of pore water flow (Naesgaard et
al. 2006).

24
Chapter 3 Soil Mixing

3 SOIL MIXING

3.1 Introduction to Mixing Phenomena

Another procedure that may result in high void ratios, excess pore water and liquefaction is
soil mixing (Naesgaard and Byrne 2005; Vasquez-Herrera and Dobry 1989). Natural water-
pluviated soils are often layered and when layers of coarse and fine grained soils are mixed
together in an undrained or near-undrained environment then the resulting post-liquefaction
shear strength of the mixed soil is much lower than the shear strength of the individual layers
(Figure 3-1 A, B, and C). There is also fourth Case D, that is not illustrated in Figure 3-1,
which results in the larger particles moving to the top during cyclic shearing (personal
communication from Professor D. Chan, University of Alberta). During the soil liquefaction
and post-liquefaction shear process there are numerous occasions during which extensive
turbulence and mixing of particles occurs. This has the potential for creating zones with
post-liquefaction strengths well below those that would be obtained from undrained
laboratory tests on the individual non-mixed layers, and is a mechanism for achieving the
strength loss required for flow liquefaction. Mixing may act independent of, and/or together
with, pore water redistribution phenomena.

Mixing may be initiated by several mechanisms including:

Pre-liquefaction mechanical mixing of fine and coarse layers such as induced


by shearing during slope failure;

Cyclic mixing where cyclic shearing induces the fine particles to fall into the
voids of an adjacent coarse layer; and,

Post-liquefaction mixing, including:


- mechanical mixing due to shear failure of liquefied soils;
- mixing due to pore water turbulence and failure at interfaces between
coarser sand and overlying finer non-plastic silt barrier layers. This may
happen upon the initiation of water interlayers or may happen when a
water bubble breaks through a confining silt layer (such as that which
occurs during the formation of sand boils).

25
Chapter 3 Soil Mixing

The ease with which cyclic mixing occurs is postulated to be a function of the
relative particle size of adjacent coarse and fine layers. If the voids in the coarse layer can
contain the particles of the fine layer, then upon being disturbed (by landslide or earthquake
shaking) may migrate into the voids of the coarse particles as illustrated in Figure 3-1 B and
C. Whereas, if the voids in the coarse layer do not freely contain the particles from the fine
layer then considerably greater turbulence and energy will be required for the soils to mix.
Mixing can produce potential net volumetric strains significantly larger than those which
would occur due to shaking of the individual homogeneous layers. This can result in either
the formation of a water film (Figure 3-1B) or in a mixed soil with near zero residual shear
strength (Figure 3-1C).

In the field, mixing has been observed resulting from static slope failure in the
Mufulira Mine tailings pond (Naesgaard and Byrne 2005). Mixing has also been observed in
videos from shake table tests with non-plastic silt barriers by Kokusho (2003) and in
simplistic plastic tube shear tests (Naesgaard and Byrne 2005).

3.2 Mufulira Mine Case History (mixing due to static slope failure)

Byrne and Beaty (1997); Barrett and Byrne (1988); and Byrne (1991), postulated soil mixing
as the reason for the flow of tailings into the Mufulira Mine in Zambia. At this site, a tailings
pond overlaid underground mine workings as shown schematically in Figure 3-2. In 1968,
the underground hanging walls of the mine collapsed and sink holes with steep side slopes
formed in the tailings pond resulting in two cases of minor inflow of tailings into the
underground workings. Following that, in 1970, a large sink hole formed and failed back
along a very shallow angle of 10 degrees (Figures 3-2 and 3-3) to develop a crest diameter of
over 250m. One million tonnes of tailings flowed into the underground mine workings,
killing 89 people (tailings.info 2005). From studies in 1988/89 by Barrett and Byrne (1988),
Byrne (1989), and Horsfield and Been (1989), it was noted that the tailings within the
overlying pond were comprised of alternating layers of sandy silt and coarser silty sand.
Whereas, the tailings within the core of the sink holes and those that had flowed into the
underground mine workings (residual in-rush or mixed tailings), contained the same
materials, but they were highly mixed and homogeneous (Figure 3-4). The silty sand made
up about 65% of the total deposit, and sandy silt the remaining 35%. The layers were clearly

26
Chapter 3 Soil Mixing

visible with thickness averaging 100 to 150mm but varying from less than 50mm to 250mm.
Water contents were in the range of 21 to 34% with an average of 25% and average in-situ
void ratio was about 0.65 to 0.7. An extensive series of undrained triaxial compression and
extension tests on the soil of the individual layers yielded high residual undrained strengths
at the in-situ void ratio and could not account for the liquefaction observed at the site.
However, when the fine and coarse materials were mixed at the same average in-situ void
ratio corresponding to an average undrained state, the resulting soil had essentially zero
strength (Figures 3-5 and 3-6). It is postulated that the sides of the 1970 sinkhole were
initially steep and then failed progressively, as illustrated in Figure 3-7. As the soil on the
initially steep-sided sinkhole failed, the silt and sand layers mixed, leading to a near complete
loss of strength that allowed the soils to flow as liquefied slurry into the underground mine
workings.

3.3 Cyclic Mixing in Rudimentary Laboratory Shear Tests

Rudimentary cyclic shear tests have been conducted in order to demonstrate whether an
earthquake or other cyclic shear loading may also induce mixing in susceptible layered soils
(Naesgaard and Byrne 2005). A clear vinyl tube of 12mm inside diameter with plugged
ends, as illustrated in Figure 3-8, was used for the tests. The samples were prepared by
removing the plug from one end, filling the tube with water, and pluviating the sand through
the water. Once a layer of sufficient thickness was achieved, the surface of the sand was
lightly tapped with a rod to render the surface level, then another layer was placed in a
similar manner, if required, or the plug was placed and clamped. The sample was then
manually sheared by grabbing the top and bottom of the tube and pushing it back and forth
while keeping the top and bottom of the sample parallel. Sample heights ranged from 27 to
40mm. Strain on typical cycles was estimated to be in the range of 10% to 20%. Table 3-1
summarizes the configurations and results of six typical tests. The grain-size distributions of
the sands and glass beads used are shown in Figure 3-9. Migration of the fine sand particles
into the voids of the underlying coarser material was readily evident during Test A. The
thickness of the water layer that formed at the top of the sample (Figure 3-10) stabilized with
approximately 20 shear cycles. When the same test as A was carried out with
homogeneous fine sand only or coarse sand only (Tests B and C), the induced

27
Chapter 3 Soil Mixing

volumetric strain was appreciably smaller (Figure 3-11). Tests D (Figure 3-12) and E
show that mixing becomes substantially more difficult as the size difference between the
coarse and fine layer decreases. Test F showed that cyclic mixing does not readily occur
without gravity or some other mechanism to move the fine materials into the coarser layer.

3.4 Discussion

In general, the susceptibility to mixing induced strength loss is deemed to be dependent on


particle and layer properties (particle shape, grain-size gradation of individual layers, layer
thickness), in-situ effective stresses, the disturbing forces (hydraulic gradient, gravity,
number and magnitude of shear strain cycles), orientation of layers relative to the disturbing
forces, saturation, and drainage constraints (permeability, presence of low permeability
barrier and rate of disturbance). Several conditions must be met before mixing, and the
resulting dramatic reduction in shear strength, will occur.

(1) Alternating layers of coarse and fine grained soil must be present. Ideally, the layer
thicknesses should be such that the available volume of fine grained soils will fit
within voids of the coarse grained layers.

(2) A disturbing force must be present to initiate the mixing. This could be a high
hydraulic gradient with associated low effective stress and turbulent flow, gravity
induced shear (slope failure), cyclic loading (from an earthquake, wind, waves or
machine vibrations), or a combination of these conditions.

(3) A substantial volume of water must be present in the voids and this water should not
be able to escape quickly so that the mixing results in low inter-particle stresses
(low effective stresses) and liquefaction, rather than drainage and consolidation.

(4) Fine particles must be able to migrate into the voids of the coarser layer. Ideally,
the fine particles should fit between the voids of the coarse layer. The ease with
which mixing occurs is proposed to be related to criteria similar to the
((D 15 ) filter /(D 85 ) soil ) < 4 to 5 criteria used to prevent internal erosion when designing
granular filters. Preliminary conclusions from the rudimentary cyclic shear tests are
that if the ratio ((D 15 ) coarse layer / (D 85 ) fine layer ) > 3 to 4, then cyclic mixing may occur,

28
Chapter 3 Soil Mixing

whereas, if the ratio is less than this, the soils will not mix easily. Cohesive and/or
cemented soils also would not mix as readily.

In the rudimentary cyclic shear tests, mixing progresses with each loading cycle
until the sample is totally mixed. The cyclic shearing action dislodges the fine particles from
the soil skeleton in the vicinity of the contact with the coarse layer, and then gravity carries
the fine particles into the voids of the coarse layer. When the coarse layer is above the fine
layer (Test F), the shearing action still dislodges the fine particles from the soil skeleton
near the interface with the coarse layer. However, in this case, there is no mechanism to
move the loosened fine particles up into the coarse layer and, therefore, mixing either does
not occur, or is minimal. It should be noted that the rudimentary shear tests carried out as
part of this research were intended to provide qualitative comparisons only. The tubes were
of relatively small diameter and wall friction and distortion of the walls would affect the
results. The amplitude of the induced shear strain per cycle was also highly variable and
relatively large (est. 10-20%). The conclusion from these tests was that cyclic shearing could
induce mixing only when specific grain size criteria of the layers were met. For normal
earthquake shaking and layering pre-liquefaction cyclic shear induced mixing is probably
minimal. Further higher quality laboratory work is warranted to better define the conditions
under which cyclic shear induces mixing.

In other situations, the fine particles may be moved by a hydraulic gradient and pore
water flow, rather than by gravity. With a sufficient flow, mixing could occur between
vertical boundaries (such as, the core of a dam), with fine particles moving into a coarse
layer. It is postulated that mixing will occur with greater ease when effective stresses are
low, as is the case when soil has liquefied. Mixing due to a hydraulic gradient and related
turbulent pore water flow generated by soil liquefaction can be observed at some of the
boundaries between coarse and fine sand and between sand and non-plastic silt in the movies
from the shake table tests by Kokusho (1999, 2003) (Figure 3-13).

Baziar and Dobry (1995), Amini and Qi (2000), have conducted cyclic triaxial tests
on layered silt-sand samples without having reported any significant mixing effects during
the tests. In both cases, the layered samples behaved similar to that of homogeneous samples
of the individual layers. The reason for mixing not having occurred in these cases may be

29
Chapter 3 Soil Mixing

due to the grain-size criteria (((D 15 ) coarse layer / (D 85 ) fine layer ) > 3 to 4) not being met, or
possibly to the presence of some plasticity, or, possibly, to the triaxial loading, and related
stress regime, not being as conducive to mixing as was the case in the elementary shear tests
carried out as part of the authors research. Remoulding (mixing) was postulated by Baziar
and Dobry (1995) and Byrne and Beaty (1997), as a possible reason for the low residual
strength observed at the end of flow failure of the Lower San Fernando Dam. Triaxial tests
by Baziar and Dobry (1995) (Figure 3-14) showed that the residual strength of the mixed soil
was 4 times less than the residual strength of the layered soil with the same void ratio.

Mixing can occur independently, or together with, the pore water migration void
redistribution mechanism. Both mechanisms may produce water interlayers and/or very low
soil strength zones. If mixing and pore water redistribution occur together, then pronounced
strength loss and near zero strengths can be expected. Mixing may result in a water
interlayer as observed in the rudimentary cyclic shear tests and illustrated in Figure 3-1 B,
or just in an extremely loose liquefied soil as illustrated in Figure 3-1 C, in the Mufulira
Mine case history, and in the liquefied zones of the Lower San Fernando Dam.

3.5 Chapter 3 Summary and Conclusions

Many natural soils and man-made hydraulically placed soils are layered. When the fine and
coarse layers mix, the resulting residual shear strength is much lower than that of the
individual layers at the same void ratio. This mixing provides explanation for the flow
liquefaction observed at the Mufulira Mine site and, at least partially, accounts for the low
end-of-failure residual shear strength back-calculated for the Lower San Fernando Dam
failure. Mixing is not a new concept, but is a mechanism that can both result in liquefaction
and lower the undrained strength of liquefied soils. The concept of cyclic mixing was
proposed by the author and rudimentary cyclic shear tests were developed by him to review
this concept.

The rudimentary cyclic shear tests were carried out within a 12mm internal diameter
vinyl tube to investigate cyclic shear induced soil mixing. The ratio of the grain-size of the
coarse and fine layers is shown to be a key factor in a layered soils susceptibility to mixing.
A preliminary postulation by the author is that cohesionless soils may be susceptible to cyclic
mixing when ((D 15 ) coarse layer / (D 85 ) fine layer ) > 3 to 4. It is also postulated that mixing will

30
Chapter 3 Soil Mixing

more readily occur when effective stresses and related inter-particle contact forces are low.
Mixing is probably more common as a post-liquefaction mechanism and helps explain the
low post-liquefaction strengths and large run-out distances observed in some instances.
Within the low effective stress low strength environment both mechanical shear and
turbulent flow can induce mixing. The occurrence of mixing induced by turbulent flow is a
concept postulated by the author following examination of the shake table test videos by
Kokusho (http//:www.civil.chuo-u.ac.jp/lab/doshitu/index.html).

31
Chapter 3 Soil Mixing

TABLES

Chapter 3
Soil Mixing

32
Chapter 3 Soil Mixing

Table 3-1 Summary of Rudimentary Cyclic Shear Tests

Post-shearing
Sample
Water (D 15 ) coarse layer (c)
Test Description (prior Volumetric Remarks
Film (D 85 ) fine layer
to shearing)(a) thickness
Strain (b)
(%)
(mm)
13mm fine sand Fine and coarse sand mixed
A over 20mm coarse 5.5 17 1.7/0.5 = 3.4 in approximately 20 shear
sand cycles (Figure 10).
Sample given 50 shear cycles
B 36 mm fine sand 0.5 1.4 NA
(Figure 11).
Sample given 50 shear cycles
C 36mm coarse sand 0.5 1.4 NA
(Figure 11).
No water layer or mixing
15mm glass beads with 20 cycles, but sample
D over 25mm coarse 0 to 3.5 0 to 9 0.8/0.5 = 1.3 mixed when given over 100
sand shear cycles and light tapping
of specimen (Figure 12).
Minimal mixing with 20
10mm fine sand cycles, but sample mixed
E over 20mm medium 1 to 4 3 to 13 1.0/0.5 = 2 when given over 100 shear
coarse (mc) sand cycles and light tapping of
specimen.
24mm coarse sand
Minimal mixing with over
F over 12mm fine 1.0 2.8 1.7/0.5 = 3.4
100 shear cycles.
sand(d)

(a)
Dimensions given are thickness of layers. Sand gradations are shown in Figure 3-9.
(b)
Volumetric strain = water film thickness/total thickness x 100
(c)
D 15 = particle diameter with 15% of particles smaller; D 85 = particle diameter with 85%
of particles smaller
(d)
Note: In this case, the coarse sand is above rather than below the fine sand.

33
Chapter 3 Soil Mixing

FIGURES

Chapter 3
Soil Mixing

34
Chapter 3 Soil Mixing

Figure 3-1 Graphic illustrating the effects of soil mixing. Elements 'A', 'B' and 'C' have the same
number and volume of large and small particles. A is the element with a fine layer
over a coarse layer prior to disturbance. 'B' represents the case where fine particles fall
into the voids between large particles and leave a water filled void or interlayer at the
top of the element. C represents the case where the particles mix evenly and the
larger particles move into a looser arrangement. Both B and C will have
significantly lower shear strength then A if mixing does not occur (Naesgaard and
Byrne 2005).

Figure 3-2 1970s Mufulira Mine tailings liquefaction failure.

35
Chapter 3 Soil Mixing

Figure 3-3 Photograph of 1970 Mufulira Mine sink hole (tailings.info 2005).

Figure 3-4 Grain-size of Mufulira Mine tailings (Barrett and Byrne 1988; Horsfield and Been
1989).

36
Chapter 3 Soil Mixing

Figure 3-5 Void ratio vs. undrained residual shear strength ( res = ( 1 - 3 )/2) relationship for sandy
silt, silty sand, and mixed residual in-rush tailings. Note: Mixed tailings have
essentially zero shear strength at the in-situ void ratio (Naesgaard and Byrne 2005 from
triaxial tests by Horsfield and Been 1989).

Figure 3-6 Conceptual undrained triaxial compression stress path for silty sand, sandy silt and
mixed silt-sand at same void ratio and initial stress state (Naesgaard and Byrne 2005).

37
Chapter 3 Soil Mixing

Figure 3-7 Postulated progressive failure mechanism of 1970 sinkhole at Mufulira Mine. (1) Sides
of initially step sided sinkhole fails, (2) as soil slides to bottom of slope, the layers mix
and due to this mixing lose nearly all strength (liquefy), (3) the liquefied soil flows
away into mine workings, (4) mechanism is repeated giving a progressive failure
mechanism, and (5) the final stable slope at shallow 10 angle (Naesgaard and Byrne
2005).

Figure 3-8 Rudimentary cyclic shear test with saturated layered sample (Naesgaard and Byrne
2005).

38
Chapter 3 Soil Mixing

Figure 3-9 Grain-size of fine and coarse sand used in rudimentary simple shear tests (Naesgaard
and Byrne 2005).

Figure 3-10 Test A before and after cyclic shearing (Naesgaard and Byrne 2005).

39
Chapter 3 Soil Mixing

Figure 3-11 Post-shearing for non-layered Tests B and C (Naesgaard and Byrne 2005).

Figure 3-12 Before and after cyclic shearing for Test D with glass beads and coarse sand
(Naesgaard and Byrne 2005).

40
Chapter 3 Soil Mixing

Figure 3-13 Slide from Kokusho cylinder shake table test video showing zone of mixing between
coarse and fine sand caused by pore fluid flow and turbulence (Kokusho 2003) (Video
from: http//:www.civil.chuo-u.ac.jp/lab/doshitu/index.html).

41
Chapter 3 Soil Mixing

Figure 3-14 Steady state void ratio vs. residual strength (S us ) for reconstituted (remolded) Lower
San Fernando Dam (California) soil. Note that the residual strength (S us ) for the
combined (mixed) homogeneous specimens is much lower than that of the layered not-
mixed specimens (Baziar and Dobry 1995). (Without mixing the strength is about
0.31 tsf whereas with mixing the strength is about 0.06 tsf).

42
Chapter 4 Residual Strength from Back Analyses of Case Histories

4 RESIDUAL STRENGTH FROM BACK-ANALYSES OF CASE


HISTORIES
4.1 Introduction

Field experience from previous earthquakes indicates that residual strengths can be much
lower than values obtained from undrained laboratory tests on undisturbed samples. This is
proposed to be due to the presence of low permeability barriers, void redistribution, soil
mixing, and stress redistribution. Based on back-analysis of field case histories, Seed (1984,
1987), Seed and Harder (1990), Olson and Stark (2002), Idriss and Boulanger (2008), among
others, have proposed residual shear strength ( res or S r ) or strength ratio (S r /p') for liquefied
soil as a function of either SPT (N 1 ) 60 or normalized CPT tip resistance Q c1cs (Figure 4.1).
The recent work by Idriss and Boulanger (2008) includes different relationships for cases
with and without void redistribution (Figure 4.1(d)).

The general procedure as described by Idriss and Boulanger (2008) involves


performing post-earthquake static limit-equilibrium slope stability analyses. The residual
strength or residual strength ratio of the liquefied portion of the failure surface is adjusted
until the calculated factor of safety is one. Non-liquefied portions of the failure surface are
assigned best-estimate shear strengths. The use of the initial (pre-failure) slope geometry
gives an upper-bound strength while use of the post-failure geometry gives a lower value.
Various procedures which account for sliding inertia and evolving geometry have been used
to interpolate between these two values. The calculated residual strength or strength
ratio is then correlated to a fines corrected SPT N-value or CPT tip resistance Q c1cs value
representative of the liquefied zone within the slope.

It should be noted that back-calculating a residual strength is not a simple task.


There are many uncertainties, unknowns, and items to consider, including:

Soil stratigraphy, density (or (N 1 ) 60 or Q c1 ) of various zones, water table at


time of the earthquake;

Time of occurrence of the flow failure (relative to earthquake shaking),


deformed shape at the time of the failure, presence of shaking and inertial
effects at time of failure;

43
Chapter 4 Residual Strength from Back Analyses of Case Histories

Extent of liquefaction and extent of the failure surface where residual


strengths are to be back-calculated;

Pore pressure regime and soil strengths to be used in the non-liquefied zones
of the back-calculation failure surface - should peak undrained strengths or
strengths with steady state pore pressures, etc. be used in the non-liquefied
zones?;

Variation of residual strength during the failure process mixing and pore
water void redistribution is likely during the failure process and strengths may
be less than the initial value when failure initiated;

Methods for using residual strengths in design. For example, if circular limit-
equilibrium analyses without any inertial effects are to be used for design,
should the same be used for back-analyses? Whereas, if the residual strengths
are to be used to calculate run-out distances of the failure, then residual
strengths back-analyzed with consideration of momentum and post-failure
geometry should be used;

Variable(s) to be used to correlate the residual strength (i.e., initial effective


stress (' vo ), fines corrected and normalized SPT N-value (N 1 ) 60cs , fines
corrected and normalized CPT tip resistance Q c1cs , static bias, shaking
duration, factor of safety against liquefaction, etc.).

As previously discussed, the residual strength back-calculated from case histories is


not a soil property, but an averaged strength achieved within the liquefied portion of the
failed slope. There is a significant variation and uncertainty in this parameter, and more than
one solution. There is also a problem in terms of similarity between the design problem
considered and the case histories upon which the correlations are based. For example, using
residual strengths back-calculated from failed embankments may not be applicable for
analyzing the lateral movement of piles through liquefied soil or using the residual strength
back-calculated from embankments with steep slopes may not be applicable for use on very
shallow slopes, etc. However, even though there are inherent uncertainties in these values,
designers have to manage with the tools and correlations at their disposal. Presently, the
back-calculated residual strength correlations may be the best information available. They

44
Chapter 4 Residual Strength from Back Analyses of Case Histories

are being used in the design of both minor and major structures. When designers use back-
calculated strengths they should endeavour to understand how they were derived and accept
that they will be working with some degree of variability and uncertainty.

Some of the key correlations that have been used in the past, or currently in use, are
briefly discussed below.

4.2 Seed (1984, 1987)

H.B. Seed back-calculated seventeen field failures and correlated the residual strength ( res )
SPT (N 1 ) 60cs (Figure 4-1(a)). For his correlations, Seed used the pre-failure geometry of the
slope or embankment, and two dimensional limit equilibrium slope analyses.

4.3 Seed and Harder (1990)

R.B. Seed and L.F. Harder repeated the back-calculated case histories carried out by H.B.
Seed but considered post-failure geometry and momentum effects (Figure 4-1(b)). They also
correlated the residual strength ( res ) to overburden corrected and fines corrected SPT N-
values (N 1 ) 60cs . The values by Seed and Harder (1990) were lower than those of Seed (1987)
due to the inclusion of momentum and evolution of slope geometry effects.

4.4 Olson and Stark (2002)

Olson and Stark updated the previous work of Stark and Mesri (1992) who argued that the
use of a residual shear strength ratio (residual shear strength divided by the initial vertical
effective stress) ( res /' vo ) should be employed rather than just res (Figure 4-1(c)). They
carried out similar back-calculations as those conducted by Seed and Harder (1990), except
that they added more data points and equated a residual shear strength stress ratio ( res /' vo )
to normalized and fines corrected SPT penetration resistance (N 1 ) 60cs . Like Seed and Harder
(1990), Olson and Stark considered momentum and the evolution of slope geometry effects.
Others who have developed ( res /' vo ) correlations include Ishihara (1993), Wride et al.
(1999), and Yoshimine et al. (1999).

4.5 Idriss and Boulanger (2008)

Idriss and Boulanger (2008) plotted selected residual strength correlations derived from
back-calculations by Seed (1987), Seed and Harder (1990), and Olson and Stark (2002) and
45
Chapter 4 Residual Strength from Back Analyses of Case Histories

then fitted suggested curves for use in design to the data (Figure 4-1(d)). Idriss and
Boulanger provide both correlations for residual strength ( res ) and residual strength ratio
( res /' vo ) versus fines corrected normalized SPT N-value (N 1 ) 60cs , and also a correlation for
residual strength ratio ( res /' vo ) versus fines corrected normalized CPT tip resistance
(Q c ) 1cs ). Suggested design curves are provided for use in situations where pore water void
redistribution would and would not occur.

4.6 Discussion

It is noted that in the above data sets the correlation between (N 1 ) 60cs (or (Q c ) 1cs )) and ( res )
or ( res /' vo ) is poor. Below a threshold, of an ((N 1 ) 60cs ) of about 12 to 15, there is little
correlation with the initial pre-liquefaction void ratio or relative density (possibly due to void
redistribution effects) and above this value residual strengths are high and flow failure
unlikely. This is a somewhat similar result to the lateral spreading estimates by Youd et al
2002 which indicate minimal displacements if (N 1 ) 60 is greater than 15.

It is also postulated that if a soil is homogeneous and expected to behave in a truly


undrained manner, then the strengths obtained from quality cyclic undrained laboratory tests
(with similar stress path and void ratio to that which would occur in the field) should be
applicable. This contention supports the use of residual strength ratios as proposed by Idriss
and Boulanger (2008), (Figure 4-1(d)) for cases without void redistribution. In Figure 4-2, a
conservative estimate of the post-liquefaction strength from undrained cyclic simple shear
tests on Fraser River Sand by Sriskandakumar (2004) (Test L14 and undrained monotonic
test) was added to the Idriss and Boulanger plot. In the simple shear test, a relative density
(D R ) of 40% (indicative of (N 1 ) 60cs 7.5 assuming D R = of ((N 1 ) 60cs /46)0.5) and initial
vertical effective stress of 100 kPa, a post-liquefaction strength of about 20 kPa was derived.

4.7 Chapter 4 Summary and Conclusions

For design, the strength of liquefied soil has been based on values back-calculated from field
case histories because alternative methodologies have not been available. Experience has
shown that strengths back-calculated from laboratory tests tend to give strengths that are too
high. This is believed to be due to the laboratory testing process not accounting for void
redistribution and mixing effects. If the in-situ soil is truly believed to behave in an

46
Chapter 4 Residual Strength from Back Analyses of Case Histories

undrained manner then the strengths from laboratory tests should be representative. In the
field, the soil is generally not truly undrained and the residual strength is dependent on
many variables (soil density, soil fabric, stratigraphy, initial stresses, duration of shaking,
permeability/groundwater flow regime, method of analyses used, etc.) and is as much a site
property as it is a soil property. Correlating the residual strength or strength ratio with soil
density or penetration resistance is a gross simplification and consequently there is much
scatter and uncertainty in the data.

The discussions in this Chapter are important in that the residual strengths back-
calculated from case histories are used in current engineering practice, and also because they
provide the input parameters for the proposed hybrid numerical procedure that is the topic of
this dissertation. In current numerical applications, the author is suggesting using a
residual strength ratio based on the charts and equations by Idriss and Boulanger (2008).
The methodology of applying these equations in the numerical modelling is discussed in
Section 6.2.

47
Chapter 4 Residual Strength from Back Analyses of Case Histories

FIGURES

Chapter 4
Residual Strength from Back Analyses of Case Histories

48
Chapter 4 Residual Strength from Back Analyses of Case Histories

Figure 4-1 Post-liquefaction residual strength form case histories (a) Seed et. al. (1984),
(b) Seed and Harder (1990), (c) Olsen and Stark (2002) and (d) Idriss and
Boulanger (2008).

49
Chapter 4 Residual Strength from Back Analyses of Case Histories

Figure 4-2 Post-liquefaction residual strength form case histories by Idriss and Boulanger (2008)
with undrained strength inferred from University of British Columbia laboratory tests
by Sriskandakumar (2004) superimposed (Byrne et al., 2008).

50
Chapter 5 Analysis Procedures Overview

5 ANALYSIS PROCEDURES OVERVIEW

5.1 Introduction

Key aspects to be addressed in the design of earth structures subjected to earthquake shaking
are (i) will liquefaction triggering occur?, (ii) is post-liquefaction stability adequate to
prevent a flow slide?, and (iii) are the induced deformations acceptable? For many years, the
state of design practice for this assessment has been based on procedures developed by Seed
and colleagues in the 1970s and 1980s (Seed and Idriss 1971; and Seed 1987). In Seeds
approach, the design problem was divided into separate analysis steps approximately as
follows:

(1) Carry out an equivalent-linear elastic dynamic analysis to determine maximum


cyclic stress ratio or earthquake loading applied to the soil;

(2) Assess whether liquefaction will be triggered using empirical correlations between
cyclic stress ratio and soil penetration resistance (N 1 ) 60 values;

(3) Assign an empirical residual strength using correlations as presented in Section 4 to


the soil zones which liquefied. Next carry out a limit equilibrium stability analysis.
If the factor of safety is less than 1.0 to 1.2, then a flow slide is assumed to occur
and remedial measures to prevent this may be required.

(4) If a factor of safety greater than 1.0 to 1.2 was calculated in (3), flow sliding was
deemed not to be a problem; however, significant deformations could still occur
during shaking. Deformations from shaking (cyclic mobility deformations) were
traditionally estimated using empirical correlations (Youd et al. 2002) or using limit
equilibrium stability analyses with Newmark-type displacement estimates
(Newmark 1965; Bray and Travasarou 2007)). If the calculated deformations were
excessive remedial measures were deemed necessary.

For many years, the standard for ground response analyses to simulate the site
effects of earthquake shaking has been to use the equivalent linear method with the computer
program SHAKE (one dimension) (Schnabel et al. 1972; Idriss and Sun 1992), FLUSH (two
dimensions) (Lysmer et al. 1975) or variants of them. The equivalent linear method is an

51
Chapter 5 Analysis Procedures Overview

iterative linear-elastic dynamic analysis process which indirectly incorporates the hysteretic
and modulus reduction with strain soil behaviour by adjusting the modulus and viscous
damping to match pre-determined modulus versus strain and damping versus strain
relationships (Figure 5-1). The equivalent-linear method has several shortcomings including:

(i) The soil will be over-damped and soft during low-amplitude cycles and under-
damped and overly stiff during large amplitude cycles;

(ii) The procedure is elastic and does not correctly simulate soil yielding or allow
irreversible permanent displacements to occur. This is especially important when
there is a static bias. Then plastic deformation will cause deformation marching
to occur in the direction of the bias;

(iii) Unnatural resonance may occur;

(iv) The pre-determined modulus versus strain and damping versus strain relationships
are based on uniform loading cycles and the correlation used to convert peak strain
from earthquake loading to equivalent uniform cycles is only approximate; and,

(v) Pore pressure generation (liquefaction triggering and consequences) is not


accounted for and must be carried out as a separate analysis (i.e., liquefaction
triggering assessment, etc.).

In recent years, the use of fundamentally more correct, generally more complex,
hysteretic non-linear analyses has become more common (DESRA (Finn et al. 1977), TARA
(Finn et al. 1986), FLAC with hysteretic damping (Itasca 2008), D-MOD2000 (Matasovic
and Ordonez 2007), UBCHYST (this dissertation), UBCSAND (Byrne et al. 2004),
SANISAND (Taiebat et al. 2010), etc.) and are replacing the equivalent-linear method. The
complex hysteretic non-linear models can be subdivided into effective stress and total
stress categories. The effective stress models account for shear-induced pore pressure
changes during the dynamic analyses, whereas, the total stress models assume constant pore
pressure during the dynamic analyses.

The current state-of-the-art/practice and the effective stress procedures proposed in


this dissertation, involve carrying out a single dynamic analysis where pore-pressure
development, potential liquefaction triggering, and resulting deformations are all determined

52
Chapter 5 Analysis Procedures Overview

in chronological sequence in a single analysis, analogous to real behaviour. Alternative


approaches are available for this single analysis. One approach is to carry out total stress
analyses where there are no changes in pore water pressure during earthquake shaking, and
liquefaction is triggered when a threshold of cycles of cyclic stress is reached. The
UBCTOT model (Beaty and Byrne 1999, Beaty 2001) is one example of this approach.
Another approach is to carry out a coupled effective stress analysis using models or programs
such as UBCSAND (Byrne et al. 2004), TARA-3 (Finn et al. 1986), CYCLIC (Para 1996;
Elgamal et al. 1999), DYNAFLOW (Prevost 2002) and SANISAND (Taiebat et al. 2010). In
the effective stress approach, the constitutive model couples shear strain with soil skeleton
volume change. If the soil is saturated then a decrease in skeleton volume (contraction)
induces a pore pressure increase and expansion of skeleton volume (dilation) induces pore
pressure decrease. Change in pore pressure results in changes in effective stress, soil
strength, and soil modulus. The UBCSAND model has been used for the work in this
dissertation, and is discussed in more detail later in this Section.

Simpler total stress constitutive models may be used for portions of the numerical
model where soil liquefaction is not deemed to occur (such as, clay, clayey silt or granular
soils above the water table). The UBCHYST model has been developed for this purpose.
The advantage of UBCHYST over even simpler elastic plastic models, such as, Mohr
Coulomb, is that it provides a more realistic replication of soil behaviour. The modulus
varies with stress ratio and damping is a direct consequence of the stress-strain loops. It does
not require the use of high Rayleigh damping values that greatly decreases run speed in the
program FLAC (ITASCA 2008), will not cause un-natural resonance and does not require
modulus versus strain relationships as part of the input.

5.2 State-of-Practice Total Stress Analysis (UBCTOT)

UBCTOT (Beaty and Byrne 1999; and Beaty 2001) addresses the liquefaction response
taking pre-triggering, triggering, and post-triggering aspects into account in a single analysis.
The earth structure and interacting structural elements are modeled in two dimensions as a
collection of discrete elements. Initially, stiff moduli, representing pre-liquefaction-
triggering conditions are specified. The FLAC Mohr Coulomb model with equivalent linear
shear modulus (G) and undrained bulk modulus (K) are used. Then, during dynamic

53
Chapter 5 Analysis Procedures Overview

shaking, the model tracks the dynamic shear stress history cyc within each element, where
cyc = st - xy ; st equals the static shear prior to dynamic excitation and xy is the shear
stress on the horizontal plane. The irregular shear stress history caused by the earthquake is
interpreted as a succession of half cycles. Each half cycle of cyclic shear stress is
transformed into an equivalent number of cycles N eq at 15 , where 15 is the cyclic shear
stress required to cause liquefaction in 15 cycles. So a small pulse may account for a portion
of a reference cycle, whereas, a large pulse may account for several reference cycles. If the
reference cycle count threshold is reached (N eq 15), then liquefaction is triggered in the
element by changing the soil strength and stiffness to the post-liquefaction values. Figure
5-2 illustrates the weighting curve used to establish the correlation between cyc and N eq .
Figure 5-3 illustrates the stress strain response used with the UBCTOT model before and
following liquefaction. Typically, Rayleigh damping of 4-8% is used with UBCTOT.

With the model, the most severely loaded or looser zones will liquefy first, the
extent of liquefaction will expand with further shaking, and stress redistribution and base-
isolation effects will occur in a somewhat similar manner to that which would occur in the
field. If sufficient elements liquefy and their residual strength is not adequate for stability,
then large deformations (flow slide) will occur. The reference cyclic stress ratio to cause
liquefaction in 15 cycles is obtained from state-of-practice liquefaction triggering charts such
as that shown in Figure 5-4 and the post-triggering strength is obtained from back-analyzed
case history residual strengths (Figure 4-1).

The procedure makes use of both state-of-practice triggering and back-analyzed


residual strength charts and is a logical extension of the current state-of-practice Seed
approach, discussed in Section 5.1. Pore pressure changes are not computed in this
approach, but are indirectly accounted for by prescribing much softer post-liquefaction
stress-strain relations and residual strength values after liquefaction has been triggered.

5.3 UBCSAND Coupled Effective Stress Analysis

5.3.1 Introduction to UBCSAND

A coupled effective stress constitutive model (UBCSAND) and analysis procedure for
modeling earthquake shaking and liquefaction has been developed at the University of

54
Chapter 5 Analysis Procedures Overview

British Columbia. UBCSAND is an elasto-plastic effective stress model with the mechanical
behaviour of the sand skeleton and pore water flow fully coupled (Beaty and Byrne 1998;
Byrne et al. 2004; and Seid-Karbasi 2008). Plastic shear induces a change in the soil
skeleton volume and, if the pores are filled with water, a change in pore-pressure and
effective stress is induced. The model includes a yield surface related to the developed
friction angle or stress ratio (Figure 5-5) and a non-associative flow rule with the soil
contractive when the stress ratio (shear stress on plane of maximum shear divided by mean
effective stress (/' m ) where (' m =' x + ' y )/2) is less than the phase transformation friction
angle and dilative when the stress ratio is greater than the phase transformation friction angle
(Figure 5-6). The soil is loading when the stress ratio increases, unloading when it decreases,
and cross-over (change from unloading on one side to loading on the other) occurs when the
shear stress on the horizontal plane ( xy ) changes sign (Figure 5-7). On unloading (and
reloading on some versions of UBCSAND) the model is elastic with bulk and shear moduli
that are a function of mean effective stress.

Increases in the mobilized stress ratio (d) is a function of plastic shear modulus
(Gp/') and the hardener (dp), as illustrated in Figure 5.8. Key elastic and plastic parameters
used are adjusted so as to give a good match with simple shear laboratory tests as the loading
path of this test, including rotation of principal stress axes, closely approximates that which
occurs during earthquake loading. The UBCSAND constitutive model is run within the
finite difference program FLAC (ITASCA 2008). In the FLAC program, dynamic analyses
are carried out in the time domain with full coupling between groundwater flow and
mechanical loading. The FLAC fish code for UBCSAND1ver2 and a flow diagram
illustrating UBCSAND are given in the attached Appendix A.

The UBCSAND model has been calibrated against simple shear laboratory tests,
centrifuge tests with and without impermeable silt barriers (Yang et al., 2004; Phillips et al.
2004; Phillips and Coulter 2005; Seid-Karbasi et al. 2005; and Park 2005) and the empirical
liquefaction triggering charts such as those by Idriss and Boulanger (2008). The model is
able to simulate both the observed drained behaviour of loose sand soils (contraction when
sheared below the cv , (or pt ) and dilative above the cv (or pt )) and the build-up of pore
pressure and soil liquefaction that occurs in undrained simple shear tests. The model is also

55
Chapter 5 Analysis Procedures Overview

able to simulate the behaviour of a flow failure when a low permeability barrier was present
and no failure when the barrier was absent or when drains were installed through the barrier
(Byrne et al. 2006; Naesgaard et al. 2005) and able to simulate the post-shaking failure of the
Lower San Fernando Dam (Naesgaard et al. 2006). Atigh and Byrne (2004), showed that
UBCSAND could simulate the behaviour of the triaxial tests with fluid inflow carried out by
Vaid and Eliadorani (1998).

A DLL version of UBCSAND (Version 904aR) and documentation regarding its


use has been uploaded to the ITASCA UDM website (Beaty and Byrne 2011).

5.3.2 Elastic Response

As indicated above, unloading and reloading (Figure 5-7) are elastic with shear (Ge) and bulk
(Be) moduli that are isotropic and a function of mean normal effective stress (') as follows:
[5-1] Ge = Ke G P a ('/P a )n

[5-2] Be = Ke B P a ('/P a )m

where:
Ge = elastic shear modulus
Ke G = shear modulus number that is a function of relative density
Often taken as 21.7 A ((N 1 ) 60 )0.33 where A = 15 to 20
or can also be back-calculated from shear wave velocity (V s ) data as
follows:
n
Ke G = V s 2 / [Pa (' / Pa) ] where = soil density

Ke B = bulk modulus number = Ke G where = function of Poissons ratio and


ranges from 0.67 to 1.33.
Pa = atmospheric pressure
' = mean normal effective stress = (' x + ' y )/2
n m 0.5

56
Chapter 5 Analysis Procedures Overview

5.3.3 Plastic Response

Plastic volumetric (p v ) and shear (p s ) strains may occur when on the yield surface. Plastic
volumetric strains are shear induced and are contractive below the phase transformation
stress ratio ( pt ) (or cv ) and dilative above, as illustrated on Figure 5-6 and by the flow rule,
Equation 5-3 below:

[5-3] d(p v )/d(p s ) = -tan() and -sin() = (sin( cv ) )

where: = dilation angle

cv = phase transformation friction angle or stress ratio


= developed stress ratio sin( f ) where f = peak friction angle

Raising of the yield surface (d) is carried out through a plastic shear modulus (Gp)
and the plastic shear strain increment (d(p s )) (the hardener) as follows:

[5-4] d = d(p s )Gp/'

The plastic shear modulus is related to and p s through a hyperbolic relationship


which for first time or virgin loading is as decribed below and illustrated in Figure 5-8.

[5-5] Gp = Gp i (1 - / f R f )2

where: Gp i = plastic modulus at =0 and equals function of (Ke G and stress level)
= developed stress ratio
f = stress ratio at failure = sin( f )
Rf = constant that truncates hyperbolic curve (between 0.7 and 1.0)

For subsequent loading, plastic strains still occur but the plastic shear modulus (Gp i )
is stiffer and increases with the number of cycles. On the other hand, following significant
dilation during the last one half cycle, Gp i is greatly reduced to compensate for the plastic
volume change (contraction) that occurs on real soil on stress ratio unloading but does not
occur in UBCSAND due to it having elastic unloading (Figure 5-9).

57
Chapter 5 Analysis Procedures Overview

5.3.4 Recent Improvements/Options added to UBCSAND

5.3.4.1 Pull Down Yield Surface on unloading if no Cross-over

In past versions of UBCSAND, unloading (Figure 5-9) and reloading (on same side) would
be elastic unless there had been a cross-over (going through zero shear stress on the
horizontal ( xy ) plane). With the proposed modification, the yield surface (Figure 5-5) is
partially pulled down (drops with the decreasing stress ratio) on unloading, even if there is no
cross-over. This results in increased plastic deformation on reloading (as the yield surface is
pushed up) and more permanent deformation in the direction of the static bias. In order to
eliminate numerical instability, a switch was added which did not allow the yield surface
pull-down to occur when effective stresses were very small.

5.3.4.2 Modeling of Dense Soils with Lower Phase Transformation Friction Angle

In previous versions of UBCSAND, the constant volume or phase transformation friction


angle was usually fixed (typically at 33 degrees) and the peak friction angle was set equal to
the phase transformation friction angle plus an increment (generally set equal to (N 1 ) 60 /
10). With the previous model, dense sands could be calibrated to trigger to liquefy at the
correct number of cycles for a given stress ratio. However, the pre- and post-triggering
response was too stiff and soft, respectively. Improved response for dense sands was
obtained by the author by reducing the phase transformation friction angle and increasing the
peak friction angle for denser sands. Suggested current correlations for use in
UBCSAND1v02 are:

[5-6] Phase transformation friction angle ( pt ) = 40.0 ((N 1 ) 60 ) 0.65) 33

[5-7] Peak friction angle ( f ) = 33 + (((N 1 ) 60 ) / 10)1.65

5.3.4.3 Delay Dilation following Post-Liquefaction Cross-over

When loose to medium dense soils liquefy, there is a period following cross-over during
which the sample strains with very little or no gain in shear strength (Figure 5-10). However,
at some limit, strain dilation kicks in, reduces pore pressure and increases strength. This
behaviour is simulated by putting in a dilation delay, with zero dilation until the cumulative

58
Chapter 5 Analysis Procedures Overview

plastic shear strain since last cross-over, is greater than the cumulative plastic shear strain of
the previous half cycle times a calibration factor.

5.3.4.4 Dilation Reduction or Cut-off upon Expansion of Element

When elements expand due to shear induced dilation or due to the flow of pore water into an
element, the dilation angle should reduce and eventually go to the critical or steady state
value of zero. This concept has been incorporated by setting the dilation angle to zero upon
the element volume change exceeding some trigger volume change. This trigger volume
change is less than that which would normally cause an element to go to the critical state
(Figure 5-11) due to localization and element size effects (Section 5.3.5), due to UBCSAND
using an elastic bulk modulus, due to numerical limitation within FLAC causing it to not
recognize a bulk modulus less than approximately 1/20th of the bulk modulus of the fluid
(ITASCA 2008) (Section 5.3.5) and due to practical limitations on the number of elements
that can be placed below an impermeable barrier (Section 5.3.5.5). In a real sand element,
the dilation would be a function of void ratio and effective confining pressure and would
decrease (less dilative) progressively as void ratio and confining pressure increase, until
reaching the critical state, at which time it would be zero. Due to the limitations discussed
above and to keep the model simple, it has been decided to cut the dilation in one step upon
reaching a prescribed volume change threshold. In UBCSAND/FLAC, this is done by
setting the negative of the sine of the dilation angle (m_dt in UBCSAND) to be greater
than or equal to zero (does not permit shear induced dilation but will allow contraction) when
the volumetric strain (vol_strain parameter in FLAC) increment since the start of
earthquake shaking reaches a prescribed limit. Due to localization and limitations in
UBCSAND and FLAC, the threshold has, to date, largely been selected by trial in one
dimensional columns and initial runs of two dimensional models. In UBCSAND analyses,
sand subjected to earthquake shaking initially has a net contraction, and then expands if there
is fluid inflow and/or continuous shear in one direction. The model seems to approximately
match model tests and field case histories when the dilation cut-off trigger volume is set
close to the original volume prior to the on-set of earthquake shaking (i.e., vol_strain0).
This gives realistic response for sand with an (N 1 ) 60 of about 10 and elements of about one
metre height.

59
Chapter 5 Analysis Procedures Overview

5.3.5 UBCSAND/FLAC Limitations and Challenges

Like all constitutive models, UBCSAND is a simplification and compromise. It is designed


for modeling saturated sand soils subjected to earthquake shaking and related potential
liquefaction triggering. The model is reasonably good at predicting skeleton volume
changes, build up in pore pressures, and number of cycles to liquefaction triggering. There
are, however, limitations within the model of which the user should be aware. Some of these
limitations are related to the constitutive model, some to the workings of FLAC, and some
due to modeling three dimensional phenomena in two dimensions. Some of the key items
are:

Low Water Compressibility

UBCSAND is, typically, run using a fluid modulus that is about one quarter that of
the actual fully saturated fluid modulus. It has been found that this value is a good
compromise between realistic modeling of liquefaction triggering and numerical time step
(proportional to run time). A portion of the modulus reduction can be justified if the actual
pore water is not fully saturated. The use of the lower modulus is partially mitigated by
using the same low fluid modulus when calibrating the model for liquefaction triggering.

Localization and Element Size Dependency

As previously mentioned, pore water void redistribution in soil profiles with low
permeability barriers can cause localization below the barrier and the analysis results become
element size dependent (Yang and Elgamal 2002; Seid-Karbasi and Byrne 2004; Naesgaard
et al. 2005; and Seid-Karbasi 2008)). Smaller (thinner) elements will give more realistic
behaviour; however, when using FLAC this is often impractical due to analysis time
constraints. Analysis time step is a function of element stiffness and minimum dimension.
Analysis time is a function of the time step and number of elements in the model. To counter
this effect, large elements can be made to behave like small elements by curtailing dilation in
the large element at a void ratio that is considerably less than that calculated from critical
state theory (Naesgaard et al. 2005; Seid-Kabasi 2008) (Figure 5-11).

60
Chapter 5 Analysis Procedures Overview

Post-liquefaction Settlement

When a sand soil is shaken, the skeleton volume contracts slightly and load is
transferred to the pore water. This process occurs with very small volume change because
water is almost incompressible. However, with time, often after end-of-shaking, the excess
pore water pressure dissipates, the effective stress increases, and the soil consolidates. As
indicated in Sections 2.1 and 2.2, if the soil has liquefied and effective stresses are zero or
near zero (little or no static bias) then much of the post-liquefaction settlement will be plastic
(i.e., non-recoverable). In current versions of UBCSAND and FLAC, there are two aspects
that lead to the calculated post-liquefaction settlement being too small (in some cases by a
factor of about 5 to 10). One problem is lack of volumetric plasticity. UBCSAND has no
volumetric yield cap and the bulk modulus is elastic. The modulus is a function of mean
normal effective stress and when the effective stress goes to zero the modulus goes to a low
default value. However, with dissipation of pore pressure, the modulus starts increasing
immediately, whereas, in real liquefied soil, a significant portion of the volumetric
deformation would be plastic and occur with no or little change in effective stress. It should
be noted that this error is mainly a problem when there is little or no static bias. As indicated
in Figure 2-5, when there is a significant static bias, sedimentation settlement does not occur
and the elastic response as assumed in UBCSAND, will be more appropriate.

In FLAC, there is also a numerical problem when the skeleton bulk modulus is
much lower (if more than approximately 20 times less) than the fluid bulk modulus. The
program will run; however, the low skeleton bulk modulus will not be recognized by FLAC
and consolidation settlements will be underestimated. One option for mitigating this is to
reduce the pore fluid modulus as the skeleton modulus decreases. This, however, is fraught
with other problems, one being that the fluid modulus is a nodal property, while the
volumetric strain is an element property. Therefore, changing modulus in one element
affects adjacent elements. The other problem, if the fluid modulus is unrealistically low, is
that unrealistic expansion of the fluid will occur when there is flow from zones with high
pressure to zones of low pressure.

Addition of a plastic volumetric yield envelope to UBCSAND may help to alleviate


some of these problems. A new version of UBCSAND with this is currently under

61
Chapter 5 Analysis Procedures Overview

development by the author. However, even if UBCSANDFLAC can be fixed to simulate


the post-liquefaction consolidation correctly, there is still a run-time issue. Post-liquefaction
consolidation can take several hours or days to occur in the real world. In dynamic analyses,
current versions of FLAC/UBCSAND, typically take approximately 15 minutes to run one
second of dynamic time. At this rate, a day of real-world consolidation would take
approximately 900 days of computer time, which is obviously impractical. In some cases,
FLACS fast-flow logic will be applicable and can be used to reduce run times to reasonable
values. Increasing the permeability throughout the model can also decrease the problem run
time, in some cases.

Multi-directional Shaking not captured in 2D Model

As in all two dimensional modeling, there is no applied out-of-plane earthquake


motion. This additional motion should accelerate the on-set of liquefaction and lead to
higher overall pore pressures for locations where liquefaction is not triggered. Stress
redistribution and pore pressure migration effects in the out-of-plane direction will also not
be detected by the model. The out-of-plane behaviour can be partially compensated for by
calibrating the triggering model to field case history triggering charts, such as that shown in
Figure 5-4; however, out-of-plane stress and pore pressure redistribution effects will not be
detected. In the future, these problems can be mitigated by adapting the model to three
dimensions and using a program, such as, FLAC 3D (ITASCA 2008).

FLAC Element Node Property Errors

FLAC stores pore pressures, fluid modulus, fluid and soil mass at the element
nodes, and then calculates element properties for use in the constitutive model by averaging
the values at the nodes. This leads to a smearing of properties at internal boundaries. For
example, when a sand layer underlying a lower permeability silt layer has an influx of pore
water and the element truly liquefies (R u =1.0) then the pore pressure at the upper boundary
of the sand layer should also go to R u =1.0. However, the pore pressure at the boundary node
will be an average value between the pore pressure in the silt and that in the sand causing R u
in the element under the barrier to be less than one. The problem can be mitigated by using
multiple smaller (thinner) elements. However, analysis speed is inversely proportional to the

62
Chapter 5 Analysis Procedures Overview

minimum elements dimension. Therefore, the modeling is often a compromise between


reasonable run times and representative soil layer dimensions.

5.3.6 UBCSAND/FLAC Grid-Model Development Considerations

Developing a UBCSAND/FLAC grid and model is a compromise between computer run


times and model accuracy. Some considerations include:

Maximum grid size for wave transmission: ITASCA 2008 recommends that the
dimension (width or height) of the element should be less than about 1/10th of the wavelength
of the highest frequency component that contains appreciable energy. For many earthquake
scenarios, this frequency can be taken as 8 to 15 hz and element sizes in the range of one/half
to two metres work well. The input record should be filtered of motions with frequencies
higher than the desired value.

Number of sand layers underlying an impermeable barrier: Seid-Karbasi 2008


suggested that there should be multiple ( 10) elements underlying a barrier in order to
properly model pore water void redistribution effects. However, unless the spacing of low
permeability layers is in the range of 5m or more this grid density is not possible. Two
options are available; use fewer elements between the barriers and accept that pore water /
void redistribution affects may not be fully captured and/or artificially space the barrier
layers further apart than they are in the field and accept that capturing one barrier correctly is
better than having multiple layers that are not captured correctly. By using an appropriate
dilation cut-off for location with expanding elements, pore water / void redistribution effects
can be captured with fewer than 10 elements underlying the barrier. The Lower San
Fernando Dam had alternative coarse and fine sand with spacing of less than the 0.5m
element size used in the back-analyses (described in Section 7.8). In order to capture the
void redistribution effects not all the silt barrier layers were included. In the lower portion of
the dam, five sand elements (approximately a combined 2.5m thickness) were placed below
the lower silt barrier. These five layers, plus an appropriate dilation cut-off allowed
localization and post-shaking failure to be captured by the UBCSAND/FLAC model.

63
Chapter 5 Analysis Procedures Overview

5.4 Hysteretic Model for Non-liquefiable Clay/Silt Soils (UBCHYST)

5.4.1 Introduction

When a soil mass is subjected to earthquake shaking, the primary loading is cyclic shearing
in the horizontal plane. The cyclic shearing induces shear stress strain behaviour that is
hysteretic in nature with characteristics as illustrated in Figures 5-12 and 5-13. These
features include:

Increasing hysteresis and reduction in secant modulus with greater strain


(Figure 5-13a);

Increasing hysteresis and reduction in secant modulus with number of cycles


(Figure 5-13b);

Permanent strains bias ratchetting when loaded with a static bias (Figure
5-13c);

Potential pore pressure generation that is a function of soil properties, cycle


amplitude and/or number of loading cycles. Increased pore pressure results in
increasing hysteresis, modulus reduction, and in the limiting condition soil
liquefaction and flow (Figure 5-13d);

Permanent secant modulus reduction with damage and pore pressure build-
up;

Strength reduction with plastic strain.

In this Section, a robust, relatively simple, total stress model, UBCHYST developed
at the University of British Columbia for dynamic analyses of soil subjected to earthquake
loading, is presented. The model is intended to be used with undrained strength
parameters in low permeability clayey and silty soils, or in highly permeable granular soils,
where excess pore water would dissipate as generated. The model has been developed for
use in the two dimensional finite difference program FLAC (ITASCA 2008) but can be
adapted to other finite difference or finite element platforms.

64
Chapter 5 Analysis Procedures Overview

5.4.2 Non-Linear Hysteretic UBCHYST Model

The essence of the proposed hysteretic model is that the tangent shear modulus (G t ) is a
function of the peak shear modulus (G max ) times a reduction factor that is a function of the
developed stress ratio and the change in stress ratio to reach failure. This function is shown
in Equation (5-8) and illustrated in Figure 5-14.

[5-8] G t = G max * (1 - ( 1 / 1f )*R f )n * Mod1 * Mod2


where: = developed stress ratio = ( xy / v )
1 = change in stress ratio ( xy / v ) since last reversal
= - max
max = maximum stress ratio () at last reversal
1f = change in stress ratio to reach failure envelope in direction of loading
= f - max
f = (sin( f ) + Cohesion * cos( f )/ v )
xy = developed shear stress in horizontal plane
v = vertical effective stress
f = peak friction angle
R f , and n = calibration parameters with default values of 1 and 2, respectively
Mod1 = a reduction factor for first-time or virgin loading (typically 0.6 to 0.8)
Mod2 = optional function to account for permanent modulus reduction with large
strain = (1- 1 / 1f )rm)*dfac 0.1 where rm and dfac are calibration
parameters

Stress reversals occur when the absolute value of the developed stress ratio () is
less than the previous value and a cross-over occurs if xy changes sign. A stress reversal
causes 1 to be reset to 0 and 1f to be re-calculated. However, the program remembers the
previous reversals ( 1old and 1fold ) so that small hysteretic loops that are subsets of larger
loops do not change the behaviour of the large loop (Figure 5-13a).

In the equation above, the tangent shear modulus varies throughout the loading
cycle to give hysteretic stress-strain loops with the characteristics illustrated in Figure 5-13.

65
Chapter 5 Analysis Procedures Overview

5.4.3 Implementation of UBCHYST Model

The equation above has been combined with the Mohr Coulomb failure criterion and
incorporated into FLAC as a user defined constitutive model (other platforms could be used
in lieu of FLAC). Typically, the model is used for fine grained soils (silts and clays) with
undrained strength parameters or with free draining granular soil with drained strength
parameters. Often it is used in combination with other models, such as, the coupled effective
stress model UBCSAND, where UBCHYST is used for the clayey non-liquefiable soils,
while UBCSAND is used for the sandy soils that are potentially liquefiable. Steps for
implementing the UBCHYST model are typically:

(1) Calibration of the model by cyclically loading a single element (this can be done
within FLAC or using a spreadsheet);

(2) Constructing a one or two dimensional model grid with appropriate soil properties.
Typically, the model is initially taken to static equilibrium in stages; initially, using
an elastic model, followed by a Mohr Coulomb model, and finally, using a
UBCHYST model;

(3) FLAC is changed to dynamic configuration and the model is brought to static
equilibrium by stepping the model with zero input motion;

(4) Earthquake motion is applied to the base of the model and dynamic analysis is
performed to the end of the record;

(5) Deformations, time histories, response spectra, etc. are obtained.

Input parameters for the model include: the maximum shear modulus (G max )
(derived from in-situ shear wave velocity measurements), bulk modulus (set equal or slightly
greater than the G max ), soil density, and Mohr Coulomb failure criterian parameters (c and
f ). A small viscous (Rayleigh) damping component (fraction of critical damping of 0.5%
to 1%) is used to give numerical stability at small strains. Calibration parameters (R f , n,
dfac, and rm) vary the shape of the stress strain loops and are adjusted so the model response
approximately matches published modulus reduction and damping data or laboratory test
results.

66
Chapter 5 Analysis Procedures Overview

5.4.4 UBCHYST Calibration

The model can be calibrated by comparing uniform cyclic response to that inferred from
published modulus reduction and damping curves similar to that shown in Figure 5-15 and/or
by comparison to the results of cyclic simple shear laboratory tests (Figure 5-16). The
simple shear test is preferred over triaxial loading because the loading path with rotation of
principal axes, etc. more closely resembles the stress path from earthquake loading.

5.4.5 Summary

A simple hysteretic total stress constitutive model (UBCHYST) has been developed at the
University of British Columbia for one and two-dimensional earthquake ground response
analyses. In the proposed hysteretic model, the shear modulus is a function of stress ratio
and varies throughout the loading cycle to give hysteretic stress-strain loops of varying
amplitude and area (damping) throughout the earthquake excitation. The magnitude of the
stress ratio is limited by a Mohr Coulomb failure envelope. The model parameters allow
calibration against laboratory data and/or published modulus-strain and damping-strain
correlations. To date, the model has been used as a user defined constitutive model within
the commercially available program FLAC. However, it could easily be adapted to other
program platforms. The model is relatively simple, robust, runs quickly and has been used
effectively on several large engineering projects.

5.5 Chapter 5 Summary and Conclusions

This Chapter introduces the constitutive models UBCSAND and UBCHYST that are key
parts (especially UBCSAND) of the proposed hybrid analysis procedure. UBCSAND when
run within the program FLAC is a coupled effective stress model that considers shear
induced soil skeleton volume changes, their effect on pore water pressure, and the resultant
pore water flow from one element to another. As indicated in Chapter 2, these are key
behaviours needed to capture soil liquefaction and related soil strength changes. The
UBCSAND program was developed by Dr. Byrne and his students. New features added to
UBCSAND as part of this research include:

Elimination of numerical instability for certain cases with static bias,

Improvement in the simulation of cases with large static bias,

67
Chapter 5 Analysis Procedures Overview

Improvement in the liquefaction behaviour by adding a dilation delay (period


during which shear does not induce volume change) thus allowing large shear
strains to develop when the normal effective stress and shear stress is near
zero, in agreement with behaviour observed in laboratory test data (Figure
5-10),

Eliminated shear induced dilation when an element reaches a critical volume


or void ratio thus making the model compatible with critical state theory and
allowing post-liquefaction flow, and

Improvements in the modelling of dense soils by making the phase


transformation stress ratio a function of soil density and changing the
correlation of peak friction angle with soil density.

In Section 5.3.5 is a list of some limitations, and challenges with the


UBCSAND/FLAC modelling is also given.

UBCHYST is a simple robust constitutive model often used together with


UBCSAND for the numerical modelling of soil structures subjected to earthquake shaking.
UBCSAND is used for portions of the soil profile deemed susceptible to liquefaction and
UBCHYST is used for those deemed not susceptible. The advantage of the UBCHYST
model over a simpler Mohr Coulomb model is the non-linear hysteretic loops developed by
varying the tangent shear modulus during loading and unloading. This more closely
replicates the behaviour of real soil and reduces the Rayleigh damping requirements; thus
increasing analysis speed. The initial algorithm for UBCHYST was developed by Professor
Byrne while the author coded UBCHYST as a constitutive model in FLAC, made
improvements regarding unload/reload loops, developed options for strength reduction as a
function of strain or number of cycles, and developed documentation explaining the model.

68
Chapter 5 Analysis Procedures Overview

FIGURES

Chapter 5
Analysis Procedures Overview

69
Chapter 5 Analysis Procedures Overview

Figure 5-1 In the equivalent linear method initial G(1) and (1) are estimated. An elastic analysis
is carried out with these values and from the results a new (1)eff is obtained. This
(1)eff is used to get a new G(2) and (2) which are used for the second elastic
analysis iteration. This process is repeated until the process converges and strain
compatible GFinal and final are obtained.

Figure 5-2 Example of the liquefaction triggering methodology in UBCTOT. The CSR at A
would cause liquefaction in 15 cycles; however, since it is only a cycle this pulse
provides 1/15 x = 1/30th of the loading to cause liquefaction. Pulse B would
cause liquefaction in 2 cycles and, therefore, provides x = of the loading
necessary to cause liquefaction. The full cycle of A and B would provide +
1/30 = 17/60 of the loading necessary to liquefy the element.

70
Chapter 5 Analysis Procedures Overview

Figure 5-3 Bilinear response of an element that liquefies when using UBCTOT (Beaty, 2001).

71
Chapter 5 Analysis Procedures Overview

Figure 5-4 Liquefaction triggering chart and equations from Idriss and Boulanger (2006, 2008).

Figure 5-5 Yield surface in UBCSAND. The hardener (d = d(/')) that expands the yield
surface (i.e., from A to B) is a function of the plastic shear strain and plastic shear
modulus as illustrated in Figure 5-8.

72
Chapter 5 Analysis Procedures Overview

Figure 5-6 Flow Rule in UBCSAND. Below cv or pt shear strain induces volumetric
contraction and above it induces volumetric dilation.

Figure 5-7 Stress ratio history showing definition of loading, unloading and reloading UBCSAND
(Beaty 2009; Beaty and Byrne 2011).

73
Chapter 5 Analysis Procedures Overview

Figure 5-8 Plastic strain increment and plastic modulus in UBCSAND.

Figure 5-9 Comparison of postulated post-liquefaction response of loose sand and that modeled by
UBCSAND.

74
Chapter 5 Analysis Procedures Overview

Figure 5-10 A dilation delay (proportional to dilation (expansive plastic volumetric strain) in last
cycle) is induced at shear stress cross-over in UBCSAND.

Figure 5-11 Large element(c) emulating the behaviour of a small element (b). If (a) and (b) are at
critical state with l/l = L/L then (c) can be made to behave similarly to (b) by setting
dilation to 0 when the volumetric strain is equal to l/L

75
Chapter 5 Analysis Procedures Overview

Figure 5-12 Loading in simple shear induces hysteretic response in soil with shear modulus (G) and
loop-size (damping) varying with strain.

Figure 5-13 Characteristic behaviour of soil in cyclic shearing.

76
Chapter 5 Analysis Procedures Overview

Figure 5-14 UBCHYST model key variables.

25
(kPa) 20 1

15 0.8
10 0.6
G/Gmax

5
0.4
0
-0.5 0 0.5 0.2
-5
(%) 0
-10
0.0001 0.001 0.01 0.1 1
-15 (%)
UBCHYST Constitutive Model
-20
Seed & Idriss (1970) Sand Upper Bound
-25
Seed & Idriss (1970) Sand Lower Bound

peak = 34.0
cohesion (kPa)= 0.0
45
'v (kPa)= 40.0 40
Damping ratio (%)

Vs (m/s)= 141.7 35
density (kg.m3)= 2041.0 30
Gmax (kPa)= 41000.0 25
n= 2.5 20
Rf= 0.8 15
10
5
0
0.0001 0.001 0.01 0.1 1
(%)

Figure 5-15 Example calibration of UBCHYST constitutive model to G/G max and damping ratio
curves.

77
Chapter 5 Analysis Procedures Overview

Figure 5-16 Comparison in response between laboratory simple shear test and UBCHYST
simulation for case with static bias (note that the scales on plots are not the same)

78
Chapter 6 Proposed Combined Effective Stress Total Stress Approach

6 PROPOSED COMBINED EFFECTIVE STRESS TOTAL


STRESS APPROACH
6.1 Introduction

As discussed in Chapter 5, fully coupled effective stress numerical analyses have been
developed (Naesgaard and Byrne 2007; Naesgaard et al. 2009). These can model the
triggering of liquefaction, the cyclic mobility deformations that occur during shaking, and
simulate the pore pressure migration and trapping of water under low permeability barriers as
observed in the field. They have also captured the behaviour observed in centrifuge test case
histories. However, there are some numerical difficulties and potential errors, including
localization-element size effects, and post-liquefaction consolidation problems, as discussed
in Section 5.3.5. There is also often limited information, and, therefore, uncertainty
regarding soil stratigraphy, soil properties, and whether key aspects are being picked up by
the numerical model. As a tool for use in current design practice, it is proposed to augment
the UBCSAND coupled effective stress analyses with a post-shaking total stress analysis
which uses liquefied (residual) strengths back-calculated from case histories. With this
approach, both the state-of-art effective stress analyses and evidence from past case histories
are considered for design. The analyses initiate as a coupled effective stress analysis that is
run up to the end of earthquake shaking. At that point, the analysis is bifurcated with one
branch being a continuation of the effective stress analysis and the other being switched to a
total stress analysis where zones that have been triggered to liquefy in the effective stress
analysis are given a residual strength derived from back-analysis of case histories. A flow
diagram illustrating the procedure is presented in Figure 6-1. If desired, post-liquefaction
consolidation settlement based on published correlations with SPT N-value ((N 1 ) 60cs ) or CPT
tip resistance (Q c1cs ) can be added to the procedure.

6.2 Implementation of Procedure in FLAC

The proposed numerical process is as follows:

(1) The model is built and brought to static equilibrium, initially using a Mohr Coulomb
constitutive model, within the program FLAC.

79
Chapter 6 Proposed Combined Effective Stress Total Stress Approach

(2) Potentially liquefiable soil zones in the model are changed to the UBCSAND
effective stress model and non-liquefiable zones are changed to the non-linear total
stress UBCHYST constitutive model. The overall model is taken to equilibrium,
first with the FLAC static configuration and then with the FLAC dynamic
configuration. Drained strengths within granular soils and undrained strengths
within fine grained soils are calculated and saved for later use. Then a dynamic
coupled effective stress analysis is carried out until the end of strong earthquake
shaking. During the analyses, zones where liquefaction has been triggered (pore
pressure ratio R u (u/' vo ) > 0.70 to 1.0) are tracked. Following the end of strong
earthquake shaking, a data file is saved (FILE A). Following this, the analyses
bifurcate and are continued as two separate cases (3a and 3b on Figure 6-1).

(3a) On Path 3a the coupled effective stress analysis is continued (until motion has
stopped and pore pressures have dissipated) to a level where flow failure from
further pore water / void redistribution is unlikely. This analysis should bring
insight into earthquake induced deformations during and following shaking and any
related localization and potential post-liquefaction flow based on the assumed
stratigraphy.

(3b) On Path 3b, the model is changed back to a total stress Mohr Coulomb model but
maintained in dynamic mode. Dilation and pore-fluid modulus are set to zero so
further shear induced pore pressure changes will not occur. The shear and bulk
modulus are left at the same value as they were in the last iteration of the previous
effective stress analysis. Elements that are deemed to have liquefied (R u > 0.70 to
1.0) during step (2) are given a residual strength that is derived from the back-
analysis of case histories and is a function of SPT (N 1 ) 60cs and possibly also vertical
effective stress. For example, using Idriss and Boulanger 2008 equation (81); if an
element has an SPT (N 1 ) 60cs-Sr = 12 and an initial vertical effective confining
pressure of 80 kPa then:

S r /' vo = exp(12/16 + ((12-16)/21.2)3 3) tan 33

= 0.105 0.65 and, therefore, S r = 0.105 x 80 kPa = 8.4 kPa

80
Chapter 6 Proposed Combined Effective Stress Total Stress Approach

where:

Sr = residual post-liquefaction strength

' vo = Initial (pre-shaking) vertical effective stress

(N 1 ) 60cs-Sr = SPT N value normalized and fines corrected in accordance with


Table 4 in Idriss and Boulanger 2008

If the element is liquefied during step (2) then the cohesion (in this dissertation,
cohesion refers to the cohesion intercept in the Mohr Coulomb model) in the
element is set to 8.4 kPa and friction angle is set to zero. If desired a factor of safety
can be applied to the residual strength by dividing it by the factor of safety.
Other, non-liquefied, elements, if fine-grained, are given their undrained shear
strengths or undrained shear strengths times a reduction factor to account for the
effects of earthquake shaking (i.e. cohesion=0.8Su and friction angle=0). Non-
liquefied granular soils are given zero cohesion and a friction angle equal to the
peak friction angle. However, if the pore-pressure in an element is less than the
initial pre-shaking static pore pressure then the elements are given an undrained
strength equal to the pre-shaking drained strength (saved at the end of step (2)). The
Path 3b analysis is continued (in dynamic mode, but with zero earthquake input)
until deformations have stopped and static equilibrium is reached. If deformations
are minimal during this phase then flow failure with traditional residual strengths is
deemed not to occur, whereas, if deformations are large, the analyses give an
indication of the potential consequences of liquefaction and flow failure. Analysis
3b runs relatively quickly and can easily be repeated with alternative residual
strength parameters.

(4) The results from both Path 3a, the coupled effective stress analysis, and Path 3b, the
combined effective stress/total stress analysis, should be considered for design.

(5) If desired, post-liquefaction consolidation settlement can be included in the model in


an approximate manner by using correlations between post-liquefaction volumetric
strain ( v ) and SPT (N 1 ) 60 such as those developed by Wu (2002), Tokimatsu and
Seed (1987), or Ishihara and Yoshimine (1992). Generally, the additional

81
Chapter 6 Proposed Combined Effective Stress Total Stress Approach

volumetric strain is only induced in elements that have liquefied. Two alternative
procedures similar to that given by Beaty (2001) are proposed. The two procedures
are similar, excepting that in one procedure, the desired volumetric strain is forced
into the liquefied element, and in the other, only the strain potential is enforced. In
an infinite slope with horizontal layers, the two procedures should give the same
volumetric strain; however, when only a local pocket liquefies, the strain potential
procedure will give less volumetric strain.

The proposed implementation is as follows:

(i) At the end of the Path 3b (or 3a) analysis (Figure 6-1) the model is changed to
a Mohr Coulomb model, as in step 3b (a file saved at the end of the step 3b
can be used). FLAC can also be changed from dynamic to static
configuration.
(ii) The volumetric strain ( v ) is then calculated for elements that are deemed to
have liquefied (R u > 0.70-1.0) using published correlations as indicated above.
An example is equation 6-1 below from Yoshimine et al. (2006) as given by
Idriss and Boulanger (2008).

[6-1] ( v ) = 1.5 exp(-0.369 ((N 1 ) 60cs )) min(0.08, max )

where:

( v ) = volumetric strain

(N 1 ) 60cs = normalized and corrected SPT N-value

max = maximum shear strain during earthquake shaking

(iii) The pore pressures in the model are set to their pre-earthquake static
equilibrium values, pore fluid modulus is set to zero to simulate drained
conditions, and shear and bulk moduli are set to the values from the last
previous iteration of the previous effective stress analysis.
(iv) A vertical stress decrement that will be used to generate the desired
volumetric strain is calculated. When a FLAC model is in static equilibrium,
changing the element modulus does not induce any strain. To get strain there

82
Chapter 6 Proposed Combined Effective Stress Total Stress Approach

has to be a change in stress. When there is a change in stress then an element


will strain in accordance with the current modulus, as FLAC is stepped back
to equilibrium. Thus, the calculated volumetric strain can be induced in the
liquefied elements by dropping the vertical total stress and letting FLAC
return (step) to equilibrium. If this is carried out in one single decrement
there is a large dynamic inertia generated and the strain will overshoot the
desired value. To minimize this dynamic effect several smaller decrements of
stress with associated stepping to equilibrium are carried out rather than one
large decrement. The vertical stress decrement to be applied to each liquefied
element is calculated as follows:

[6-2] V = ( v ) M / I $switch

where:
M = constrained modulus = K + 4/3 G
K = elastic bulk modulus
G = elastic shear modulus
v = desired volumetric strain in the liquefied element according to
published correlation
V = vertical stress in FLAC model (compression is negative)
V = decrement of vertical stress to apply in FLAC (a positive number
because compression is negative in FLAC)
I = number of iterations to be applied in FLAC as necessary to
minimize dynamic inertia (i.e. between 10 and 100).
$switch = a conditional value that is either 1 or 0. It is one when either (i)
the volumetric strain has not reached the desired v or (ii) when
number of steps completed is less than I. If the preceding
conditions are met $switch is set to zero. Condition (i) is used if it
is desired to enforce a specific strain amount into all the liquefied
elements, and condition (ii) is used if it is only desired to enforce
the strain potential into the elements.

83
Chapter 6 Proposed Combined Effective Stress Total Stress Approach

(v) A numerical loop consisting of applying the vertical stress decrement,


stepping to equilibrium, and checking the $switch condition, is repeated until
all values of $switch are zero. At that point, the desired volumetric strain or
volumetric strain potential, depending on the condition used for $switch, will
be induced in all elements that were deemed to have liquefied.

6.3 Discussion

As indicated in Section 6.1, the combined effective stress total stress analysis is undertaken
due to the possibility that the effective-stress analysis alone may not capture the full
response. The total-stress analysis on its own also is fraught with uncertainty; however, it is
tied to actual field case histories and is accepted by the engineering community as current
design practice. When the two methods are combined, the envelope of results should be
more conservative and chances of capturing the correct response improved.

In the authors opinion, the coupled effective stress dynamic analyses is the state-of-
the-art and for slopes subjected to earthquake shaking and, generally, provides the most
reliable estimates of deformation available. The procedure also provides much insight into
potential mechanisms. When the earthquake record is provided and actual stratigraphy is as
indicated in the model, then the predicted deformation (for cases where there is no flow
failure) is probably within a range of one-half to double the actual deformation. Class A
predictions of centrifuge test deformations have been in this order. For cases with flow
failure, the deformation may not be as reliable mainly due to bad geometry modelling
problems. Adding the post-shaking with residual strength in liquefied zones adds some
confidence to the results and it is suggested that this check should always be carried out. For
cases where the consequences of flow failure are severe, it may be prudent for the designer to
also consider the probability of getting a lower bound of the case history residual strength
data. One option is to progressively reduce the residual strength or strength ratio used in
the analyses until failure occurs (or the residual strength goes to zero). If failure occurs
then one needs to assess whether the probability of getting that value is acceptable.

It should be noted that the proposed total stress residual soil strengths were
originally developed using limit equilibrium slope stability methods and would not
necessarily be the same had they been calculated using the numerical methods as proposed.

84
Chapter 6 Proposed Combined Effective Stress Total Stress Approach

The differences would be mainly due to (i) deformations during shaking that may reduce the
driving stresses that may cause failure in the post-shaking total stress analysis, and (ii) pore
water pressures in non-liquefied zones may differ from the assumptions made when the
residual strengths were derived. However, there is such large scatter or variation in the back-
calculated strength parameters that the differences are not deemed to be significant. If
deformations during shaking are very large, then consideration could be given to using the
pre-earthquake geometry for the post-liquefaction residual strength analysis.

The simplified empirical post-liquefaction consolidation strains are believed to be


representative of level ground conditions and the calculated settlements likely over-estimate
settlements for conditions where there is a static bias. Incorporating these in the analysis
process as discussed in Section 6.2, Item (5) for cases where there is a significant static bias
(a relatively common situation) is expected to give conservatively large post-liquefaction
consolidation settlements. Forcing the total post-liquefaction strain ( v ) into each liquefied
element is expected to give a conservative estimate of post-liquefaction differential
settlement. Enforcing only the strain potential may give more realistic differential
settlements. However, when using only strain potential, the resulting differential settlements
are sensitive to the chosen soil modulus and strength of the liquefied soil, and care must be
taken in choosing these parameters.

Selecting elements deemed to have liquefied by monitoring the pore pressure ratio
(R u = u dyn /' vo ) is not always elementary. Sometimes, there are pore pressure spikes that
will cause an excessively high R u for a very short duration that are not representative of the
general state of pore pressure within the element, and, sometimes, there are large
deformations that cause the initial vertical effective stress (' vo ) not to be representative of
the current depth and stress state of the element. To mitigate these problems, a dual check,
using both the maximum pore pressure ratio during shaking (R u(max) ) and the pore pressure
ratio at the end of the earthquake shaking (R u(current) ) is used. Elements having both R u(max)
and R u(current) above specific thresholds are deemed to have liquefied. Typically, the R u(max)
threshold is set between 0.7 and 0.9 and the and R u(current) is set between 0.5 and 0.7.

85
Chapter 6 Proposed Combined Effective Stress Total Stress Approach

6.4 Chapter 6 Summary and Conclusions

Dynamic numerical analyses using the UBCSAND constitutive model as described in


Chapter 5 may be the state-of-the-art in predicting liquefaction induced deformations.
However, as indicated in Section 5.3.5 there are still shortcomings. In addition to the
modelling shortcomings, there is also the difficulty of characterizing the soil stratigraphy
with accuracy at any given real site. As indicated in Chapter 4, the use of these case history
derived post-liquefaction strength is current engineering practice, even though there are large
uncertainties in the back-calculated strengths. The reason the relatively inaccurate field
derived strengths are used is because there are no easy to use readily available alternatives.
When considering all this, it was proposed that a good design should consider both
approaches, and the hybrid effective stress total stress design approach described in this
Chapter is the result. A description of the procedure and step-by-step details on
implementing it are given in this Chapter. A limitation on the UBCSAND model is that bulk
modulus is elastic without any volumetric yielding cap. This results in an underestimate of
post-liquefaction consolidation settlement. To compensate for this, a procedure for
implementing empirically derived post-liquefaction consolidation settlements is also given in
this Chapter.

An example analysis illustrating the application of the procedure is given in


Appendix B.

86
Chapter 6 Proposed Combined Effective Stress Total Stress Approach

FIGURES

Chapter 6
Proposed Combined Effective Stress Total Stress Approach

87
Chapter 6 Proposed Combined Effective Stress Total Stress Approach

Figure 6-1 Flow chart showing combined coupled effective stress / total stress analysis procedure
(Naesgaard and Byrne 2007).

88
Chapter 7 Model Calibration, Validation and Application Example

7 MODEL CALIBRATION, VALIDATION and APPLICATION


EXAMPLE
7.1 Introduction

An important aspect of numerical modeling is calibrating and verifying the response of the
model. The UBCSAND effective stress constitutive model has been calibrated both to
replicate the behaviour of undrained simple shear laboratory tests and replicate the
liquefaction triggering response implied by the case-history derived liquefaction triggering
design charts. The model and proposed analysis methodology has also been verified by
comparing the numerical results against those of triaxial tests with inflow (Atigh and Byrne
2004), shake table tests, centrifuge tests, and field case histories.

7.2 Calibration to Laboratory Cyclic Simple Shear Tests

7.2.1 Introduction

A key part of developing the UBCSAND model has been making it simulate the behaviour
observed in monotonic and cyclic simple shear tests. This is both for the purpose of
verifying that UBCSAND can simulate the element behaviour observed in the laboratory and
for calibrating post-liquefaction triggering response for use in design.

As part of the research of this dissertation, calibrations have been carried out to
show that UBCSAND can simulate the behaviour observed in the laboratory testing program
by Sriskandakumar (2004). These tests were part of the UBC Liquefaction Initiative
(NSERC Grant 246394), a research program of laboratory and centrifuge testing and
numerical modeling carried out jointly by the University of British Columbia and C-CORE,
Newfoundland. The Sriskandakumar (2004) test program included undrained monotonic,
undrained cyclic, and drained cyclic tests on loose sands (Relative Density (D r = 38-44%)
and dense sands (Relative Density 80-81%). The tests were carried out on an NGI-type
simple shear device. The same Fraser River Sand and air pluviation preparation method
being used in the centrifuge tests at C-CORE, was used for the laboratory testing program.
Fraser River sand is fluvial sand deposited near the mouth of the Fraser River. The grains
are angular to sub-rounded and are composed of approximately 40% quartz, 11% feldspar,

89
Chapter 7 Model Calibration, Validation and Application Example

and remainder other minerals and rock fragments (Sriskandakumar 2004). The natural sand
typically has approximately 4-10% silt content; however, the fines were removed for the
laboratory and centrifuge testing. Figure 7-1 shows a typical grain size curve of the tested
material.

Numerical modeling of select tests from the work by Sriskandakumar (2004) was
undertaken using the UBCSAND1ver2 constitutive model. A FLAC fish code for the
model is in the attached Appendix. A single element was numerically exercised with the
same loading path as the laboratory test. The results were then compared, calibration
parameters within the numerical model were adjusted, and the process repeated until a
reasonable match between the laboratory and numerical response was achieved.

7.2.2 Results and Discussion

Figures 7-2 to 7-10 illustrate comparisons between the numerical and laboratory tests. Table
7-1 summarizes the properties, calibration factors, and cycles to liquefaction from the
laboratory tests and numerical analyses. The calibration factors were generally fixed for tests
with similar initial conditions, and factors were only changed when known initial conditions
changed. It is important to note that calibration factors can be a function of or vary with
known initial conditions, such as, initial confining pressure or initial relative density but
should not be functions of parameters (such as, cyclic stress ratio) that are not known at the
on-set of earthquake shaking. As shown on Figure 7-7 and Table 7-1, the UBCSAND model
(as calibrated) tended to trigger liquefaction later than the laboratory tests at high cyclic
stress ratios and trigger earlier at lower stress ratio. Increasing the initial elastic shear
modulus parameter m_kge improved the ability to calibrate the model to match the
laboratory test results at a range of cyclic stress ratio.

It is noted in Figure 7-7 that, for a given stress ratio, liquefaction resistance
increases with confining pressure. This is believed to be due to stress densification causing
the higher relative density with increase in confining pressure. This seems to override the k
effect which typically would cause liquefaction resistance (CRR) to decrease with increasing
confining pressure. When calibrating to the laboratory data with varying confining pressures
the use of a k factor is not required as the UBCSAND/FLAC model and laboratory test are
using the same boundary stresses and relative density.

90
Chapter 7 Model Calibration, Validation and Application Example

Figures 7-5 and 7-6 show the L12 loose sand sample with a significant initial static
bias. If the same calibration that was used for the tests without initial static bias is used for
this test the results are as indicated on Figure 7-5 and liquefaction is triggered too early;
however, the post-triggering response is reasonable. Adding an additional calibration factor
(m_hfac7) which adjusts the pull-down of the yield envelope prior to liquefaction triggering)
in addition to m_hfac6 which adjusts the pull-down of the yield envelope following
liquefaction triggering, allowed a better match to the data as shown in Figure 7-6. The
model pore pressure response and amount of permanent strain per cycle (marching) is very
sensitive to the amount of yield envelope pull-down allowed. Thus, while earlier versions of
UBCSAND, which do not allow for pull-down of the yield envelope, would under-predict
pore pressure build-up when there was no cross-over, the current UBCSAND1v02, which
pulls down the yield envelope on unloading and then generates significant pore pressures on
reloading may trigger liquefaction prior to that indicated by the laboratory tests.

Figures 7-8 to 7-10 shows the simulation by the current model of a cyclic simple
shear test on dense sand. Earlier versions of UBCSAND would tend to be too stiff prior to
liquefaction triggering and too soft following triggering. Modifications to the model,
including increasing peak friction angle, lowering of the phase transformation friction angle,
dilation delay following post-liquefaction cross-over, as described in Sections 5.3.4.1,
5.3.4.2, and 5.3.4.3, allows the model to cyclically soften gradually as observed in the
laboratory tests.

7.3 Calibration to Seed type Empirical Liquefaction Triggering Charts

Determining liquefaction triggering response from site specific laboratory tests is often not
possible or impractical due to sample disturbance problems. It is common practice,
therefore, to rely on empirical Seed type (Seed et al 1984; NCEER 1997; Cetin et al. 2004;
Idriss and Boulanger 2008) empirical liquefaction triggering charts and in-situ testing such as
Standard Penetration Test N-value or cone penetration test (CPT) tip bearing (Q c1cs ) values
when assessing liquefaction triggering. This is also the recommended procedure for use with
UBCSAND; however, for post-triggering behaviour and behaviour associated with a static
bias, reliance is still placed on laboratory tests.

91
Chapter 7 Model Calibration, Validation and Application Example

In the calibration process, a single undrained soil element is exercised so as to


trigger liquefaction in the correct number of cycles and to give post-liquefaction stress-strain
behaviour consistent with that observed in laboratory simple shear tests. The recommended
calibration procedure is as follows:

(1) Set up the 2D FLAC profile and bring it to static equilibrium. Generally, there will
be a range of soil types, confining pressures, k o values, static bias, and soil densities
(density is often represented by penetration resistance of in-situ tests). The model is
then broken into zones with similar properties (effective stresses, G max and (N 1 ) 60 ,
Q t , etc.) and representative soil elements from the cohesionless zones are then
selected from each zone for calibration. The vertical and horizontal effective
confining pressure, small strain shear modulus, and density (generally, normalized
and fines corrected SPT N-value (N 1 ) 60-CS or normalized and fines corrected CPT
tip resistance (Q c1cs )) are recorded for each element to be calibrated.

(2) An undrained single element model is set up in FLAC and is initialized with the
representative vertical and horizontal effective confining pressure, small strain shear
modulus, and (N 1 ) 60-CS .

(3) A cyclic shear stress ( xy ) compatible with a cyclic resistance ratio (CRR) that will
liquefy (pore pressure ratio (R u ) near 1.0 or shear strain ( = 3.75% )) in 15 cycles
(CRR 15 ), from an empirical liquefaction triggering chart, is then applied to the
element to be calibrated. For example, the chart and equations by Idriss and
Boulanger, 2008 as shown in Figure 5-4 may be used and xy calculated as follows:

xy = vo ' * CRR 15 * k
where:
xy = applied cyclic shear stress
vo ' = vertical effective stress
CRR 15 = cyclic resistance ratio from chart or equation
k = correction factor for confining pressure (Idriss and Boulanger
2008)
(4) The single element is then repeatedly cyclically loaded with the xy from step (3)
and calibration parameters are adjusted until the element liquefies in 15 cycles and

92
Chapter 7 Model Calibration, Validation and Application Example

the post-liquefaction stress-strain cycles are compatible with typical laboratory


simple shear tests. This process is then repeated for all the elements to be calibrated
and a matrix of calibration parameters (m_hfac1 to m_hfac4 or m_hfac6) for the
model is obtained.

(5) The calibration parameters or equations representing the parameters (as functions of
effective confining pressure, penetration resistance, etc.) are then introduced as
material parameters within the applicable zones in the larger 2D model and the
dynamic analysis continued.

Figure 7-11 compares UBCSAND element data to empirical design triggering charts
from calibrations by Beaty (2009), Beaty and Byrne (2011).

7.4 One-Dimensional Tests for demonstrating Void Redistribution

Infinite slope one-dimensional numerical analyses are useful for developing insights into the
behaviour of the low permeability barrier and flow slide mechanisms. Figure 7-12
(Naesgaard et al. 2005) illustrates a typical one dimensional column analysis with typical
volumetric strain time histories at various locations within the column. Figure 7-13 shows a
displaced grid with velocity time histories above and below the barrier. Note how a flow
slide or flow failure condition is initiated (increasing velocity) at time x-x. This is the
critical state when shear induced dilation goes to zero. Shear strain is concentrated
(localized) immediately below the low permeability barrier, and the flow failure is
independent of the inertial forces from strong shaking. Seid-Karbasi (2008) conducted a
detail numerical study of the infinite slope (one dimensional) column of sand soil underlying
a barrier. Seid-Karbasi (2008) noted that the upper approximately one-third thickness of the
sand layer (Zone E Figure 7-12) tended to be expansive due to inflow, whereas, the soil
below this was contractive (Zone C Figure 7-12). However, as discussed in Section 2.2, for
level ground conditions this may not be the case and the whole soil column may be
contractive excepting for the water layer directly under the barrier. Seid-Karbasi (2008)
showed by analyses that drainage columns installed through the low permeability barrier
were an effective mitigation measure against post-liquefaction flow failure. However, even
though the drains prevented flow failure, some movements still occurred during strong

93
Chapter 7 Model Calibration, Validation and Application Example

shaking. Taiebat (2008) also was able to generate post-liquefaction post-shaking


deformations in a 1-D columns using the effective stress program SANISAND.

7.5 George Massey Tunnel Centrifuge Test Modeling

Dynamic numerical analyses were used for the seismic retrofit design of the forty four year-
old immersed-tube George Massey Tunnel. The 1.3 km long tunnel carries four lanes of
traffic under the Fraser River just south of Vancouver, British Columbia. Design criteria
were that the retrofitted tunnel should withstand both a 0.25g magnitude 7.0 non-subduction
earthquake, and a 0.15g magnitude 8.2 distant subduction earthquake without collapse or loss
of life, but with possible damage to a repairable level including controllable water leakage.

Soil liquefaction, its consequences, and mitigation were the key design challenges in
this project. Two-dimensional dynamic analyses using the program FLAC were a prime
geotechnical analyses and design tool. Displacements from the numerical analyses were
used as input into three-dimensional static structural analyses using non-linear soil springs
and nonlinear moment-curvature section properties. The structural analyses were used to
assess and mitigate potential cracking in the tunnel.

In the 2D FLAC analyses, transverse and longitudinal sections were studied using
total and effective stress constitutive models (UBCTOT and UBCSAND) developed at the
University of British Columbia. Dynamic shaking, liquefaction triggering, consequences of
liquefaction and soil-structure interaction were addressed in each of the models. Analyses
were carried out with and without retrofit measures. A centrifuge-testing program was
carried out to check the numerical model. Pre-test (Class A) predictions of the centrifuge
test showed good agreement between the FLAC/UBCSAND numerical model and centrifuge
test results. The geotechnical design considerations, numerical modeling and centrifuge test
program are discussed in detail in Yang et al. (2005), Naesgaard et al (2004), Yang et al,
(2003), and Adalier et al (2003). Figure 7-14 shows a transverse cross-section of the
approximately one kilometre long tunnel, Figure 7-15 shows an example numerical mesh,
Figures 7-16 and 7-17 show calculated displacement vectors from the FLAC/UBCSAND
model for sloped river bottom and level river bottom, respectively. Figure 7-18 shows the
layout of the centrifuge model. Centrifuge analyses were carried out for three scenarios.
One scenario was with no ground improvement around the tunnel, the other was with 10m

94
Chapter 7 Model Calibration, Validation and Application Example

wide bands of densification on either side of the tunnel, as indicated in Figure 7-18, and
another was with 10m wide drainage zones (zones with same density as remainder of the
sand but higher permeability. Measured displacements and accelerations are compared to
numerical predictions in Table 7-2 and in Figures 7-19 to 7-20. The numerical analyses of
the centrifuge model were carried out in prototype scale.

7.6 UBC/C-CORE Centrifuge Test Modeling

A series of eight centrifuge tests were carried out at the C-CORE facility in St. Johns,
Newfoundland (Phillips and Coulter 2005; Phillips et al 2004). The centrifuge tests modeled
submerged slope configurations with and without: low permeability silt barrier, soil
densification dyke, and drainage trenches. Air pluviated Fraser River sand with a relative
density of approximately 40% and minimum and maximum void ratio of 0.62 and 0.94 was
used (Figure 7-1). Non-plastic commercial ground silica silt was used for the low
permeability barrier and clear uniform coarse sand was used for the drainage layers. D 10 and
D 50 were 0.16mm and 0.26mm for the loose sand, 0.005mm and 0.016mm for the silt, and
2.2mm and 2.9 mm for the drainage sand, respectively. The centrifuge tests were at 70g with
a water plus hydroxypropyl-methylcellulose fluid with a viscosity of 35 times that of water.
Simulated earthquake motion was applied during flight using a hydraulically actuated shaker.
All dimensions, time histories, etc. given in this section are in the scaled prototype
dimensions rather than the actual centrifuge dimensions. During centrifuge spin-up there are
large changes in effective stress which result in stress densification of loose sandy soils
(Park and Byrne 2004). This was accounted for in the analyses. A typical grid and input
earthquake record is shown in prototype scale in Figure 7-22. The side forces that would
occur within the centrifuge box were accounted for during the dynamic analysis by applying
internal nodal forces that were a function of the out-of-plane effective stress ( z ) times the
sidewall friction coefficient. Normalized velocities were used to indicate the direction of the
internal nodal forces.

Liquefaction flow failure was observed in the tests which included the low
permeability silt barrier and higher levels of shaking. Flow failure, generally, did not occur
when the barrier was absent or if drainage trenches were placed through the barrier. Figure
7-23 compares centrifuge and numerical results for the COSTA-C test (Phillips and Coulter

95
Chapter 7 Model Calibration, Validation and Application Example

2004) that included a low permeability barrier. In the COSTA-C test, a flow slide occurred
at the barrier interface at approximately 50s after end of strong shaking. Figure 7-24 shows
displaced profiles for a similar model, CT5 (Phillips et al 2004) that had permeable drainage
slots through the silt barrier. With drainage slots (Figures 7-24 and 7-25) all deformation
occurred during strong shaking (t < 20s) and flow deformation is prevented.

The response of some of the centrifuge tests was indicative of the centrifuge model
not being fully saturated. This was a quandary as P-wave measurements within the
centrifuge model were indicative of full saturation. To investigate this, a series of triaxial
tests with P-wave measurement apparatus were carried out at UBC. This testing program is
discussed in Chapter 8.

7.7 Kokusho Shake Table Emulation

Kokusho (1999), Kokusho (2000), Kokusho and Kabasawa (2003), carried out a series of
shake table tests on both cylinders of sand (Figure 2-6) and on shake table model soil
embankments for the purpose of studying the effects of low permeability barriers and pore
water migration on slope stability. Figure 7-26 shows a photograph of one of the tests (from
videos at http//:www.civil.chuo-u.ac.jp/lab/doshitu/index.html). The rectangular Lucite
shake table soil box was 1100mm long, 600mm wide and 800mm high. The sand was placed
at a relative density of approximately 27-34% and the model was shaken in the transverse
direction only to minimize inertial forces on the slope. Two thin horizontal non-plastic silt
barriers were placed in the slope as shown in the figure (Kokusho and Kabasawa, 2003).
Shaking induced pore water pressure build-up within the slope and mild movements as
shown in the end-of-shaking profile shown in Figure 7-27; however, at this stage, the model
appeared stable. Then, approximately four seconds after end-of-shaking, a flow failure
initiated with the effects of pore water interacting with the silt barriers clearly visible. Figure
7-28 shows the post-failure profile.

A UBCSAND/FLAC model of a similar test to the shake table test was set up to
demonstrate that the numerical models could simulate the failure mechanisms observed in
the shake table tests. The FLAC model slope was 16m high instead of 600mm to minimize
problems with effective stress defaults in the UBCSAND program. The model mesh with
two barriers is shown in Figure 7-29. Dynamic analyses were carried out for three conditions

96
Chapter 7 Model Calibration, Validation and Application Example

(1) with low permeability layers or barriers and pore water flow allowed, (2) with low
permeability barriers but no flow allowed, and (3) with no barrier but flow allowed. Soil
properties used in the numerical model are summarized in Table 7-3. Figure 7-30 shows the
deformed grid at the end of the analyses (30 s) for the case with barrier and flow. The
localization of shearing under the barrier is clearly visible. Figures 7-31 shows horizontal
movement contours for the case with barrier and flow and case with barrier and flow not
allowed. The effects of pore water redistribution (flow) and the barrier are clear and also
illustrated in Figure 7-32 which shows displacement time histories for the three cases
modeled. For all cases, similar deformation occurs during shaking; however, only the case
with the barrier and pore water flow allowed deforms significantly after end of shaking. A
key parameter in getting the model to flow after end-of-shaking is the dilation cut-off that is
added to the UBCSAND to simulate the behaviour when a local zone reaches the critical. In
these simulations, dilation was cut-off when the volume of any given element expanded
beyond its original volume prior to start of shaking. When this was triggered the dilation
was set to a value that was contractive to simulate some mixing occurring between the sand
and silt barrier after the soil reaches the critical state.

7.8 Lower San Fernando Dam Example Analysis

7.8.1 Introduction

In 1971, the Lower San Fernando Dam (LSFD) was shaken by a large earthquake with a
peak velocity pulse of around 0.6m/s and peak ground acceleration of approximately 0.5 g
(Seed 1973; Seed et al. 1989; Castro et al. 1989; and Castro 1995). Approximately 20 to
30 seconds after the end of the earthquake shaking, the upstream face of the dam failed
catastrophically, leaving only 1.5m of freeboard and placing a large population at risk.
Extensive investigation and analyses of the dam were conducted following the event. It was
concluded from the studies that the hydraulic fill soil within the lower and central portions of
the dam had liquefied and overlying portions of the dam had flowed out riding on the
liquefied soil (Figure 7-33).

In the current example, the liquefiable portions of hydraulic fill soils were given
(N 1 ) 60 values of 12 while the clayey core of the dam was given undrained shear strength of
18% of the initial vertical overburden pressure (Figure 7-34). The UBCSAND constitutive

97
Chapter 7 Model Calibration, Validation and Application Example

model was used for the potentially liquefiable sandy soil portions of the dam while a Mohr
Coulomb model was used for the clayey core and portions of the dam above the water table.
The effective stress analyses were fully-coupled (mechanical - pore water flow) and run
within the program FLAC.

In the actual dam, the spacing of the low permeability layers is much finer than in
the numerical model. Limitations on smallest element size are necessary for analysis time to
be in a practical range and the behaviour of finely spaced layers, therefore, must be simulated
with larger elements by increasing the distance between silt barrier layers and by introducing
a dilation cut-off in any elements expanding beyond their original volume.

7.8.2 Coupled Effective Stress Analysis of LSFD Phases 2 and 3a

In the example LSFD analysis, the coupled effective stress analysis was carried out using the
UBCSAND constitutive model within the program FLAC as described in Section 5.3 for all
saturated sand elements while a total stress Mohr Coulomb model was used for the clay/silt
core portion of the dam.

Initially, the whole model was brought to static equilibrium using a Mohr Coulomb
model and drained stiffness parameters (Phase 1). Then prior to earthquake shaking, the
sand elements were changed to the UBCSAND model while portions of the clay/silt core
were left with a Mohr Coulomb model but given undrained stiffness parameters. The model
was then brought to equilibrium in dynamic mode and all displacements reset to zero (end of
Phase 1). Following this, the dynamic analysis with earthquake motion input at the base was
run until excessive grid distortion stopped the analysis ( 40 s). This constitutes Phase 2 and
3a in Figure 6-1. A file was saved at after the end of strong shaking (15 s) for use in the
Phase 3b Total Stress Alternate Analysis.

7.8.3 Post-Liquefaction Total Stress Analysis of LSFD Phase 3b

Elements which had liquefied (deemed to occur when pore pressure ratio (R u ) max > 0.7))
during the strong shaking portion of the analysis (Phase 2) where changed to a Mohr
Coulomb model with zero friction angle, zero dilation, and a cohesion set equal to a residual
strength (S r ) calculated using the initial pre-shaking effective stress (') and the S r /' ratio

98
Chapter 7 Model Calibration, Validation and Application Example

recommended by Idriss and Boulanger (2007) (Figure 4-1(d)). The dynamic analysis was
then continued until stopped by excessive grid distortion ( 18 s).

7.8.4 Results of LSFD Example Analysis

During strong shaking (Phase 2), the upstream shell moved approximately 3m upstream and
the downstream shell moved approximately 1.5m downstream. Liquefaction occurred both
within the upstream and downstream portions of the dam as shown in Figure 7-35. Strong
shaking ended at approximately 10 seconds (Figure 7-36).

Continuing the coupled effective stress analysis (Phase 3a in Figure 6-1) resulted in
very little movement within the dam until approximately 32 seconds at which time the
upstream shell proceeded to fail in a similar manner to that reported for the actual dam
(Figure 7-36). Figure 7-37 shows displacement and velocity of the dam at 40 seconds at
which time the program stopped due to excessive distortion of some of the elements. The
downstream face did not move significantly following end of strong shaking as was the case
for the actual dam.

In the Phase 3b total stress alternative, the dam proceeded to fail once the liquefied
elements where changed to a Mohr Coulomb model with residual strength S r as given by the
Idriss and Boulanger (2008) relationship. This was not unexpected as the Lower San
Fernando Dam was one of the key case histories used in deriving the relationship. Again, the
dam only failed in the upstream direction as was the case with the actual dam.

99
Chapter 7 Model Calibration, Validation and Application Example

TABLES

Chapter 7
Model Calibration, Validation and Application Example

100
Chapter 7 Model Calibration, Validation and Application Example

Table 7-1 Summary of Undrained Laboratory Simple Shear and UBCSAND Simulations

TEST 'vo DR CSR ST/'vo NLIQ UBCSAND1v02 CALIBRATION PARAMETERS


No. LAB UBCSAND hfac1 hfac2 hfac3 hfac4 hfac5 hfac6
kPa (%) (-) (-) (-) (-) (-) (-) (-) (-) (-) (-)
Note 1 Note 2 Note 3 Note 4 Note 5 Note 6
L1 100 40 0.08 0 17.0 16.0 0.65 0.85 1.00 0.60 1.00 0.95
L2 100 40 0.1 0 6.5 7.5 0.65 0.85 1.00 0.60 1.00 0.95
L3 102 40 0.12 0 2.5 4.0 0.65 0.85 1.00 0.60 1.00 0.95
L4 100 40 0.15 0 1.0 2.0 0.65 0.85 1.00 0.60 1.00 0.95
L5 200 44 0.08 0 33.0 15.0 0.40 0.85 1.00 0.60 1.00 0.95
L6 200 44 0.1 0 7.5 7.0 0.40 0.85 1.00 0.60 1.00 0.95
L7 200 44 0.12 0 3.5 4.0 0.40 0.85 1.00 0.60 1.00 0.95
L8 200 44 0.15 0 1.0 2.0 0.40 0.85 1.00 0.60 1.00 0.95
L9 50 38 0.08 0 12.0 12.0 0.30 0.85 1.00 0.60 1.00 0.95
L10 50 38 0.1 0 3.0 6.0 0.30 0.85 1.00 0.60 1.00 0.95
L11 50 38 0.12 0 1.5 3.5 0.30 0.85 1.00 0.60 1.00 0.95
L12 100 40 0.065 0.1 15.5 4.0 0.65 0.85 1.00 0.60 1.00 0.95
L12c 100 40 0.065 0.1 15.5 12.0 0.65 0.85 1.00 0.60 1.00 0.95
D3 100 80 0.35 0 8 8 0.05 0.20 1.80 3.00 0.60 1.00

Notes:
1- Test numbers refer to tests by Sriskandakumar (2004), Chapter 4. "L" signifies loose and "D" signifies dense samples.
2- Initial effective vertical stress
3- DR= Relative density
4- CSR=Ratio of single amplitude shear stress to the initial effective vertical stress
5- TST=Static Shear Bias
6- NLIQ= Number of cycles required to reach pore pressure ratio near 1 for loose samples or number of cycles required
to reach 3.75% shear strain for dense samples.
7- For test L12c and additional hfac7 was added to control pretriggering pull-down of yield envelope

Table 7-2 Massey Tunnel Project. Class A Numerical Predictions compared to Centrifuge
Test Results (in prototype scale)

Scaling Factor
Output Parameter Model 1(1) Model 2(2) Model 3(3)
(model:prototype)
Peak tunnel heave (m) 1:100 0.25 (0.27) 0.13 (0.14) 0.12 (0.04)
Peak tunnel lateral movement (m) 1:100 0.59 (0.68) 0.50 (0.35) 0.40 (0.30)
Maximum soil displacement (m) 1:100 1.52 (1.50) 1.20 (1.30) 1.03 (1.10)
Peak tunnel horizontal acceleration (g) 100:1 0.10 (0.11) 0.13 (0.08) 0.095 (0.095)

Note: Numbers given in brackets are the results from the centrifuge tests.
(1) No ground improvement loose sand around tunnel
(2) 10m wide densification on each side of tunnel
(3) 10m wide drainage zone on each side of tunnel

101
Chapter 7 Model Calibration, Validation and Application Example

Table 7.3 Soil Properties used in simulation of Kokusho Shake Table Tests
Bulk
Constitutive Dry cv f cohesion (N1)60cs shear modulus modulus Porosity Permeability Remarks
model Density Kge Gmax G B n k
(tonnes/m3) (kPa) blows/ft (kPa) (kPa) (kPa) cm/s

SAND UBCSAND 1.5 33 34 0 10 15 (1) - 2 Gmax 0.45 5x10-3 (3) (4)

SILT BARRIER MOHR COULOMB 1.45 - 25 0 12 12.5 - (2) 2G 0.55 1x10-7

(1) Gmax = 21.7 x Kge x ((N1) 60cs)0.33 x Pa x ('m/Pa) 0.5 where Pa = 100 kPa and 'm = mean normal effective stress in kPa
(2) G = (21.7 x Kge x ((N1) 60cs)0.33 x Pa x ('m/Pa) 0.5) / 5
(3) UBCSAND calibration factors for SAND layers: m_hfac1 = 0.5, m_hfac2 = 1.0, m_hfac3 = 1.0 (used UBCSAND VERSION 15oo.FIS)
(4) Dilation cut-off in UBCSAND model: When incremental volumetric strain since start of shaking = 0.0 dilation m_dt set 0.001 gave 4m displ & if set to 0.25 gave 7m disp
(contractive dilation angles used to account for mixing)

102
Chapter 7 Model Calibration, Validation and Application Example

FIGURES

Chapter 7
Model Calibration, Validation and Application Example

103
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-1 Typical grain-size distribution and microscope view of Fraser River Sand used for
cyclic shear tests (from Sriskandakumar 2004).

104
Chapter 7 Model Calibration, Validation and Application Example

0.8 Lab data


UBCSAND1v02

u/VC'
0.6

0.4

0.2 (a)
0
0 2 4 6 8 10 12 14 16 18 20
NO OF CYCLES

(b)

30
'vc=100kPa; Drc=40%
Lab data cyc/'vc=0.08; st/'vc =0.0
20
FLAC
Shear Stress, (kPa)

10

0
-15 -10 -5 0 5 10 15
-10

-20 (c)

-30
Shear Strain, (%)

Figure 7-2 Comparison of undrained simple shear test L1 laboratory test to UBCSAND
simulation (a) pore pressure ratio vs. number of cycles, (b) stress path, and (c) stress
strain response.

105
Chapter 7 Model Calibration, Validation and Application Example

0.8 Lab data

u/VC'
0.6 UBCSANDv02
0.4

0.2

0
0 2 4 6 8 10
NO OF CYCLES

30
'vc=100kPa; Drc=40%
Lab data
cyc/'vc=0.10; st/'vc =0.0
UBCSAND1v02 20
Shear Stress, (kPa)

10

0
-15 -10 -5 0 5 10 15
-10

-20

-30
Shear Strain, (%)

Figure 7-3 Comparison of undrained simple shear test L2 laboratory test to UBCSAND simulation
(a) pore pressure ratio vs. number of cycles, (b) stress path, and (c) stress strain
response.

106
Chapter 7 Model Calibration, Validation and Application Example

0.8
(a)
u/VC'
0.6

0.4
Lab data
0.2
UBCSAND1v02

0
0 1 2 3 4 5 6
NO OF CYCLES

(b)
30
Lab data 'vc=102kPa; Drc=40%
cyc/'vc=0.12; st/'vc =0.0
UBCSAND1v02
20
Shear Stress, (kPa)

10

0
-15 -10 -5 0 5 10 15

(c)
-10

-20

-30
Shear Strain, (%)

Figure 7-4 Comparison of undrained simple shear test L2 laboratory test to UBCSAND simulation
(a) pore pressure ratio vs. number of cycles, (b) stress path, and (c) stress strain
response.

107
Chapter 7 Model Calibration, Validation and Application Example

0.8

0.6

u/VC'
0.4 Series1

0.2 (a) UBCSAND1v02

0
0 5 10 15 20 25 30
NO OF CYCLES
30
Lab data 'vc=100kPa; Drc=40%
UBCSAND1v02 cyc/'vc=0.065; st/'vc =0.1
20
Shear Stress, (kPa)

10

0
0 25 50 75 100 125

-10 Vertical Effective Stress, 'v (kPa)

-20 (b)
-30
30
'vc=100kPa; Drc=40%
cyc/'vc=0.065; st/'vc =0.1
20
Shear Stress, (kPa)

10

0
-15 -10 -5 0 5 10 15
-10

(c)
-20 Lab data
UBCSAND1v02

-30
Shear Strain, (%)

Figure 7-5 Comparison of Undrained simple shear test L12 laboratory test with initial static bias
to UBCSAND simulation (a) pore pressure ratio vs. number of cycles, (b) stress path,
and (c) stress strain response using same calibration parameters as those used without
static bias.

108
Chapter 7 Model Calibration, Validation and Application Example

0.8
(a)
0.6

u/VC'
0.4 Series1

0.2 UBCSAND1v02

0
0 5 10 15 20 25 30
NO OF CYCLES
30
Lab data 'vc=100kPa; Drc=40%
UBCSAND1v02 cyc/'vc=0.065; st/'vc =0.1
20
Shear Stress, (kPa)

10

0
0 25 50 75 100 125

-10 Vertical Effective Stress, 'v (kPa)

-20 (b)
-30 30
'vc=100kPa; Drc=40%
cyc/'vc=0.065; st/'vc =0.1
20
Shear Stress, (kPa)

10

0
-15 -10 -5 0 5 10 15
-10

(c) -20 Lab data


UBCSAND1v02

-30
Shear Strain, (%)

Figure 7-6 Comparison of undrained simple shear test L12 laboratory test with initial static bias to
UBCSAND simulation (a) pore pressure ratio vs. number of cycles, (b) stress path, and
(c) stress strain response using additional hfac7 calibration parameter to control pull-
down prior to triggering.

109
Chapter 7 Model Calibration, Validation and Application Example

0.16
Lab sig'vo=50 kPa, DR=38%
0.15
UBCSAND sig'vo= 50 kPa, DR=38%
0.14 Lab sig'vo=100 kPa, DR=40%
Cyclic Stress Ratio (CSR)

0.13 UBCSAND sig'vo=100 kPa, DR=40%


Lab sig'vo=200 kPa, DR=44%
0.12
UBCSAND sig'vo=200 kPa, DR=44%
0.11
0.1
0.09
0.08
0.07
0.06
0 5 10 15 20 25 30 35
Number of cycles to Liquefaction, N LIQ

Figure 7-7 Comparison of cycles to liquefaction between undrained laboratory simple shear tests
and UBCSAND simulations.

110
Chapter 7 Model Calibration, Validation and Application Example

60
'vc=100kPa; Drc=80%
cyc/'vc=0.35; st/'vc =0.0
Shear Stress, (kPa) 40 (a)
20

0
-15 -10 -5 0 5 10 15
-20

Point of =3.75% (i.e. Assumed


-40 triggering point of liquefaction for
comparison purposes)

-60
Shear Strain, (%)

(b)

Figure 7-8 Undrained dense sand simple shear stress vs. strain comparison of D3 laboratory test to
UBCSAND simulation (a) Laboratory test, and (b) UBCSAND simulation.

111
Chapter 7 Model Calibration, Validation and Application Example

60
'vc=100kPa; Drc=80%
cyc/'vc=0.35; st/'vc =0.0
40
Shear Stress, (kPa)

20

0
0 25 50 75 100 125
-20

-40
Point of =3.75% (i.e. Assumed triggering point of
liquefaction for comparison purposes)
-60
Vertical Effective Stress, 'v (kPa)

60
'vc=100kPa; Drc=80%
cyc/'vc=0.35; st/'vc =0.0 UBCSAND1v02
40
Shear Stress, (kPa)

20

0
0 25 50 75 100 125
-20

-40

-60
Vertical Effective Stress, 'v (kPa)

Figure 7-9 Undrained dense sand simple shear stress path comparison of D3 laboratory test (a) to
UBCSAND simulation (b).

112
Chapter 7 Model Calibration, Validation and Application Example

1
0.8
u/VC'
0.6
0.4
0.2 Lab data
0
0 5 10 15 20
NO OF CYCLES

1
0.8
0.6
u/VC'

0.4
0.2 UBCSAND1v02
0
0 5 10 15 20
NO OF CYCLES

Figure 7-10 Undrained dense sand simple shear pore pressure ratio comparison of D3 laboratory
test (a) to UBCSAND simulation (b).

113
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-11 Empirical liquefaction triggering chart with overlain UBCSAND simulations (from
Beaty 2009; Beaty and Byrne 2011).

114
Chapter 7 Model Calibration, Validation and Application Example


Figure 7-12 Volumetric strain (v) within an
infinite slope column with a low
permeability crust over loose sand.

Figure 7-13 UBCSAND numerical model of infinite slope (1D) column with
low permeability barrier cap. At time x-x' zone 'I' has expanded
to the critical state, dilation goes to zero, and flow failure is
initiated.

115
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-14 Section through Massey Tunnel in Fraser River.

Figure 7-15 Typical transverse FLAC model grid used from Massey Tunnel. Scale: the tunnel is
24m wide - vertical and horizontal scale are the same.

Figure 7-16 Typical distorted mesh with displacement vectors. Scale: the tunnel is 24m wide -
vertical and horizontal scale are the same.

116
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-17 Typical distorted mesh with displacement vectors for level ground condition. Scale:
the tunnel is 24m wide - vertical and horizontal scale are the same.

Figure 7-18 Layout of Massey project centrifuge model showing location of instrumentation Model
2 (Legend: a3 = accelerometer; P12 = piezometer; L v and L h = LVDT). Dimensions
in centimeters. Model 1 and Model 3 are similar.

117
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-19 Model #1 accelerometer A10 (dark line is FLAC class A prediction and light line is
from centrifuge test).

Figure 7-20 Comparison of class A numerical prediction and centrifuge results for Model #2 with
10m wide densification.

118
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-21 Comparison of pore pressure in piezometer P5 in Model #3 (dark colour is FLAC class
A prediction and lighter colour is centrifuge test result).

Figure 7-22 Typical grid and input time history used for UBC Liquefaction Initiative C-CORE
centrifuge tests (shown in prototype scale).

119
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-23 Comparison of Centrifuge tests and numerical results for profile with low permeability
silt barrier (COSTA-C): (a) displaced grid, (b) Horizontal displacement of sliding
block over barrier (for the centrifuge data the solid line is measured and the dashed line
is corrected to better match final displacements), (c) vertical displacement at crest, (d)
centrifuge and numerical surface profiles, (e) calculated lateral displacement contours
in metres, (f) and (g) pore pressure time histories at P3 and P6, (h) acceleration time
history at A6.

120
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-24 (a) Initial and displaced profile of centrifuge test CT5 with three drainage slots (b)
numerical analysis of same.

121
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-25 Comparison of vertical displacement near crest with (CT5) and without (COSTA-C)
drainage slots. Post-shaking flow initiated in the COSTA-C test at approximately 70s.

Figure 7-26 Kokusho shake table model at start of test (Kokusho 2003).

122
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-27 Kokusho shake table test profile at end-of-shaking.

Figure 7-28 Post-failure profile of Kokusho shake table test. Failure occurred a short time after all
shaking had stopped.

123
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-29 UBCSAND/FLAC model grid of Kokusho shake table test.

Figure 7-30 Original boundary of model and displaced shape of FLAC model grid.

124
Chapter 7 Model Calibration, Validation and Application Example

(a)

(b)

Figure 7-31 Contours of horizontal displacement (a) with barrier layers in place and flow on and (b)
with barriers but no flow.

125
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-32 Calculated displacement time histories of shake table model with and without barrier
and with and without flow.

126
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-33 Lower San Fernando Dam which failed into the reservoir approximately 30s after end
of earthquake shaking. Blue areas are soil that was mixed and deemed to have
liquefied. Lower figure is reconstructed profile which shows location of liquefied soil
before failure. (Seed et al. 1973).

127
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-34 Grid, locations of low permeability barriers, (N1) 60 and shear strength input parameters
used in Lower San Fernando numerical example.

128
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-35 Profile showing zones deemed to have liquefied at end of strong shaking (R u > 0.7).

Figure 7-36 Displacement time histories of point near upstream toe of dam for both 3a coupled
effective stress analysis (ESA) and 3b total stress analysis (TSA) alternative. Both
analyses stopped due to excessive element distortion.

129
Chapter 7 Model Calibration, Validation and Application Example

Figure 7-37 (a) Horizontal velocity (m/s) and (b) horizontal displacement (m) at end of analyses
(40s) for coupled effective stress analysis.

130
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

8 P-WAVE MEASUREMENTS AS AN INDICATOR OF


SATURATION IN CENTRIFUGE TESTS
8.1 Introduction

It has been shown that the triggering of liquefaction in sandy soils is sensitive to the
compressibility or saturation of the pore fluid (Chaney 1978; Kokusho 2000; Tsukamoto et
al. 2002; Yang 2002; Yang et al. 2004; Yang 2005). Cycles to liquefaction increase with
increasing pore fluid compressibility or decreasing saturation. Pure water is relatively
incompressible (Fluid bulk modulus 2 x 10 Pa); however, air is highly compressible and
9

the inclusion of small bubbles of air within the pore fluid (saturation < 100%) greatly
increases the compressibility of the fluid. Measuring the fluid compressibility or saturation
directly is often difficult. However, in the laboratory the compressibility and degree of
saturation of the soil can be correlated with the Skempton B (Skempton 1954) as indicated
in Equation (8-1) (Yang 2002).

1
[8-1] B=
Kb K
1+ n + n b (1 Sr )
Kw Pa

where: B = Skempton B = u/ m in undrained triaxial tests


u = change in pore pressure from an increment in mean normal pressure m
n = porosity
K b = bulk modulus of soil skeleton (individual soil grains are assumed
incompressible)
K w = bulk modulus of pore fluid
P a = atmospheric pressure
S r = saturation (as a fraction)

The B-value is easily measured in undrained triaxial tests. However, in centrifuge


and shake table tests, measuring the B-value and related degree of saturation is problematic.
It has been proposed by Yang and Sato (1998), Kokusho (2000), Tsukamoto et al. (2002),
and Yang (2002) that the P-wave or compression wave velocity (Vp) within the soil medium
can be correlated to the degree of saturation and B-value as indicated by Equation (8-2)
(Yang 2002) and illustrated in Figure 8-1.

131
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

1/ 2
4G / 3 + K b /( 1 B )
[8-2] V p =

where: G = shear modulus and = density

Data from Tsukamoto et. al. (2002), and Takahashi, et. al. (2006) showed a
correlation between laboratory measurements and V p calculated from Equations (8-1) and
(8-2) while Ishihara et al. (1998) and Tamura et al. (2002) showed that some laboratory P-
wave B-value measurements deviated from the relationship given by Equation (8-2). V p has
been used to locate zones of reduced saturation in the field (Sato and Yang 2000; Kokusho
2000) and has been correlated to Skempton B and to cyclic resistance to liquefaction in
laboratory tests (Tsukamoto et al. 2002; Ishihara et al. 1998) (Figure 8-2).

In a recent UBC Liquefaction Initiative centrifuge test program carried out to study
liquefaction-induced slope failure mechanisms, there was concern that the soil within the
model was not fully saturated and that this may affect liquefaction triggering and model
behaviour. The centrifuge model was prepared by air-pluviation of the fine to medium sand
through a funnel with the same width as the model and with the sand drop height adjusted to
achieve the desired relative density. The centrifuge test model was then inundated by:
placing the specimen under a vacuum, flushing it with carbon dioxide, and slowly
introducing a viscous pore fluid. A viscous pore fluid with a viscosity of 35 times the
viscosity of water was used to improve pore fluid flow scaling when subjected to the 70 g
centrifuge acceleration. P-wave velocity (V p ) measurements were made after inundation and
following spin-up. However, the results were noted to be erratic. In order to investigate
these anomalies in a fundamental manner, a laboratory test program was devised that would
simulate the centrifuge saturation procedure and allow the verification of saturation by
measurements of the Skempton B, and V p . Shear wave velocity (V s ) was also measured.
Numerical analyses were conducted using the program FLAC (ITASCA 2002) in fully-
coupled mechanical strain - groundwater mode in order to gain insight into the behaviour
observed. This Chapter presents the results of the laboratory program along with those from
numerical analyses seeking explanations for the erratic behaviour observed during centrifuge
modeling.

132
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

8.2 Laboratory Element Testing Program

The laboratory element testing program was designed to simulate the sand placement and
saturation procedures used in the previously mentioned centrifuge tests. The sand material
(Fraser River Sand) and pore fluid for element testing were also similar to those used in the
centrifuge tests. Figure 8-3 illustrates the laboratory setup. As indicated in Table 8-1, the
testing involved measurement of B-value, V s , and V p on two triaxial specimens (Tests A and
B) of 75mm diameter and 156mm height, under manifold testing conditions (i.e., with and
without pore fluid, with and without cell water, with varying mean normal effective stress
levels, and before and after 4- to 6-day saturation periods). All mean normal effective stress
( m ) variations on the triaxial specimens were exerted as hydrostatic loadings. Bender
elements, as supplied by GDS Instruments Ltd. (2002), were located in the centre of the top
and bottom platens; one set being the transmitter, and the other the receiver. The bender
elements could be excited either in bending to generate shear waves or axially to generate a
compression wave. Excitation frequencies were in the range of 3,500-7,000 Hz.

The Fraser River sand used for the tests is a fine to medium grained sand with D50
of 0.26mm, D10 of 0.18mm and minimum and maximum void ratios of 0.62 and 0.94,
respectively. The pore fluid used for the tests consisted of a 1.75% by weight methyl
cellulose (methocel) water solution. The methocel solution has a viscosity of approximately
35 times that of water and is used in centrifuge tests to achieve better pore fluid flow scaling.
The solution was pre-mixed using warm water. It was then cooled and de-aired by placing it
under a 50 kPa vacuum for more than 8 hours.

Specimen preparation and testing procedure were as follows:

(i) The test specimens were prepared by air pluviation of the Fraser River sand through
a funnel with an opening of approximately 2mm by 12mm. The sand was allowed
to free fall 10 to 15mm as the funnel was manually moved back and forth over the
specimen. The as-placed relative density of the specimens was approximately
20%.

(ii) The specimen was then placed in the triaxial cell. Prior to the introduction of pore
fluid, isotropic confining pressures ranging from 15 to 50 kPa were applied by
various combinations of cell pressure and pore vacuum. Both air and water were

133
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

used to provide cell pressure. P-wave and shear wave velocities were measured for
each load case.

(iii) The specimen was then placed under a pore vacuum of 40 to 50 kPa and purged
with carbon dioxide (CO 2 ).

(iv) Methocel solution (pore fluid) was introduced at the bottom of the specimen with
1.5m head under a nominal cell pressure and 50 kPa pore vacuum, as illustrated in
Figure 8-3. This was continued until clear methocel solution was exiting at the top
of specimen (2.5 to 4 hours). Following this, the specimen was cycled with
alternating 50 kPa vacuum and the methocel at atmospheric pressure. P-wave and
shear wave velocities were measured at various stages during the inundation
process.

(v) V s , V p and B were measured at various cell and pore pressure increments.

(vi) The specimen for Test A was then held under a constant back-pressure of 226 kPa
for 4 days to provide an opportunity for improved saturation; similarly, the
specimen for Test B was held under a constant back-pressure of 156 kPa for six
days.

(vii) After this holding period (for improved saturation), V s , V p and B were again
measured at various cell and pore pressure increments.

As noted in Table 8-1, cell pressure was supplied using air or water. Air was used
for portions of the test to confirm that the P-wave travel path was not through the cell fluid in
lieu of the specimen as desired. Results from the laboratory tests are summarized in Table 8-
1. As may be noted, there was no noticeable difference in the P-wave velocity when air was
used in the cell instead of water.

Figure 8-4 shows a plot of V s with respect to the mean effective stress m giving a
best fit correlation as follows:

V s = 1.7 x P a ( m / P a )0.23 [8-3]

V p for unsaturated dry specimen conditions varied between 1.90 and 1.96 V s . This
corresponds to a small strain dry sand Poissons ratio of 0.32 and a Bulk Modulus (K b ) to
Shear modulus (G) ratio (= K b /G) of 2.4.
134
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

Figures 8-5(a) and 8-5(b) show typical V p and V s time history traces for a B of 0.48
and 0.95. Figure 8-6 shows the measured V p and V s as a function of Skempton B. Note that
the arrival time does not change significantly with a change in B once the specimen is
inundated with pore fluid and that the measured V p is indicative of a near fully saturated
specimen, even for B less than 0.5. This is contrary to the findings of Tsukamoto et al.
(2002), Yang (2002), and Kokusho (2000) and correlation from Equation (8-2). As shown in
Figure 8-9, the measurements also show greater disparity from the correlation inferred from
Equation (8-2) than the data measured by Ishihara et al. (1998) and Tamura et al. (2002). As
expected, V s is essentially independent of B.

8.3 Numerical Modeling

The experimentally observed V p vs. B correlation, as illustrated in Figure 8-6 is contrary to


the expected behaviour inferred from Equation (8-2). A numerical analysis using the
program FLAC (Cundall and Board 1988, ITASCA 2002) in coupled mechanical strain
groundwater flow mode was conducted to provide insight to the behaviour observed. FLAC
is a finite difference numerical program which solves problems dynamically in accordance
with the equations of motion using an explicit approach involving very small time steps. The
formulation in FLAC is done within a framework of Biot theory with Darcy flow in a porous
medium. The approximately 75mm diameter by 150mm high specimen was modeled in
plane strain as a rectangular grid of 35 by 75 elements (Figure 8-7). The soil was modeled
with an elastic skeleton with soil and pore fluid properties as indicated in Table 8-2. The
permeability of the soil with the methocel pore fluid is 1/35th of the permeability of the same
soil with water.
The numerical modeling was carried out in two ways. In the first type of analysis,
the pore fluid was assumed homogeneous and fluid modulus (K w ) was kept at a constant
value at all nodes. This value was changed for each FLAC run to simulate varying degrees
of saturation. This analysis is deemed to be representative of the air being uniformly
disseminated within the pore fluid with many small bubbles and has been designated
homogeneous-partial-saturation (HPS). Figure 8-8(a) illustrates the HPS conditions. The
HPS type analyses produced the V p vs. B relationship shown as squares in Figure 8-9 and is
in close agreement with Equation (8-2), as illustrated by the solid line in Figure 8-9.

135
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

In the other type of analysis, the pore fluid was modeled with scattered pockets of
air bubbles. This has been designated non-homogeneous-partial-saturation (NHPS), and is
illustrated in Figure 8-8(b). For the NHPS case, the air bubbles are modeled individually as a
few pockets of large bubbles scattered throughout the grid. In the FLAC model, the air
bubbles were modeled by reducing the fluid modulus (K w ) to a low value expected within
the air bubbles (K w = P a + u) where P a = atmospheric pressure and u = pore pressure) at
selected nodes scattered throughout the grid (Figure 8-10). The fluid modulus at other nodes
was set to a high value of that representative of near fully saturated water (2.2x109 Pa). The
V p vs B correlation from the NHPS type analyses is presented in Figure 8-9 (circle symbols)
and shows a similar behaviour to that observed in the laboratory tests (triangle symbols).
The P-wave time histories from the numerical analyses were similar to those observed from
the bender elements in the laboratory.
A major portion of the numerical modeling was conducted in plane strain instead of
the axisymmetric modeling commonly used to simulate triaxial tests, in order to facilitate all
elements being the same size or volume. This was important in the NHPS analyses as it
allowed easy control of air bubble size and spacing. The HPS case was modeled both in
plane strain and axisymmetric and the results were similar. The intent of the numerical
modeling was not to give an exact replication of the laboratory model but to provide insight
into the observed behaviour.

8.4 Discussion

It is postulated that in the laboratory test there were scattered zones of bubbles surrounded by
zones with either no or minimal air bubbles (non-homogeneous-partial-saturation (NHPS)).
This will result in a stiff matrix with local soft pockets. The first arrival P-wave will be that
which routes through the stiffer zones. This behaviour was simulated in the NHPS numerical
model which showed that scattered discrete bubbles did not reduce V p even though they
dramatically reduced the Skempton B. However, if there are numerous small bubbles within
each pore space (homogeneous-partial-saturation (HPS) then the stiffness of the fluid is also
homogeneous and the P-wave velocity will be a function of Skempton B, with saturation and
fluid stiffness in accordance with Equations 8-1 and 8-2.

136
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

The NHPS saturation observed in the laboratory specimen may be partially related
to the use of viscous methocel solution in lieu of water. The high viscosity may lead to the
existence of a few local pockets of air bubbles rather than finely scattered small bubbles.
Tests by others who were using water as the pore fluid (Tsukamoto et al., 2002) have
indicated P-wave velocities in accordance with Equations 8-1 and 8-2. It is postulated that it
may be possible to achieve better saturation of the specimen by inundating the specimen with
de-aired water prior to introducing the methocel solution. A drawback of this procedure is
that pockets of water may get trapped within the specimen and not become displaced by the
methocel. It would be difficult to verify that all of the water had been displaced.

Placing the specimen under a backpressure of greater than 150 kPa and holding it
for several days was effective in increasing the Skempton B to levels indicative of good
saturation (B > 0.85). The Skempton B remained high when the backpressure within the
specimen was reduced to atmospheric pressure. It is suggested that this common laboratory
procedure be considered as part of the process for saturating centrifuge models.

8.5 Chapter 8 Summary and Conclusions

Centrifuge tests were carried out as part of the UBC Liquefaction Research Initiative
(NSERC Grant 246394) for the purpose of validating the UBCSAND numerical model.
During this work it was noted that the behaviour in the centrifuge did not match that
predicted by the numerical models and it was suspected that the pore fluids in the centrifuge
were not fully saturated, even though P-wave velocity (V p ) measurements were indicative of
full saturation. To check why this may be the case, triaxial tests that simulated the centrifuge
saturation procedure were carried out. The triaxial cells had bender elements in the top and
bottom platens that allowed measurement of shear wave velocity (V s ) and P-wave velocity
(V p ) during the testing. The results from the testing showed that when the measured
Skempton B in the triaxial sample was high (representative of full saturation), the
corresponding V p was also high. However, a high V p did not always correspond to a high
Skempton B. It is postulated that the behaviour in the later case is due to the air bubbles
giving the low B-value being widely spaced, thus, allowing the P-waves to find travel paths
through the medium that bypassed the bubbles (thus giving the high velocity). From this,
two types of partial saturation have been defined: homogeneous-partial-saturation (HPS)

137
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

whereby the air bubbles are numerous, very small and included within each void space
throughout the specimen, and non-homogeneous-partial-saturation (NHPS) whereby the air
bubbles are much larger and are only present within some pore spaces. For HPS, the V p will
be a function of the Skempton B and average saturation (S r ) av , whereas, for NHPS, V p is
independent of Skempton B and average saturation. With NHPS, high V p values indicative
of high saturation are possible, even though the B value is relatively low. NHPS conditions
are postulated to be more prevalent in centrifuge tests where viscous solutions, such as
methocel, are often used in lieu of water. In addition to the laboratory test, the behaviour
was also numerically simulated using the program FLAC. This later simulation provided
confidence in the postulated explanation for the behaviour.

When the triaxial specimen was placed under backpressure for several days this
resulted in air going into solution and increased saturation and related Skempton B values.
This procedure is standard practice in the saturation of laboratory specimens, and it is
proposed that the procedure also be considered for saturating centrifuge test specimens.

138
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

TABLES

Chapter 8
P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

139
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

Table 8-1 Summary of Laboratory Test Conditions and Results

Initial Initial Final Final


Mean Shear
Cell Pore Cell Pore
Effective P-wave Wave Pore Cell
Skempton
Pressure Pressure Pressure Pressure
Stress (1) Velocity Velocity fluid (2) Fluid
'B'
(P i ) (u i ) (P f ) (u f )
(' m ) f (V p ) (V s )
(kPa) (kPa) (kPa) (kPa)
(kPa) (m/s (m/s)
TEST A
--- --- 0 -15 15 --- 240 123 air air
--- --- 15 -10 25 --- 252 142 air water
TEST A SPECIMEN PURGED WITH CO2
50 10 99 27 72 0.34 (0.36) 1420 159 meth. water
99 27 152 44 108 0.32 (0.36) 1420 177 meth. water
152 44 198 60 138 0.35 (0.36) 1562 184 meth. water
198 60 298 109 189 0.49 (0.50) 1562 200 meth. water
298 109 362 146 216 0.58 (0.59) 1562 208 meth. water
358 113 398 137 261 0.59 (0.61) 1562 226 meth. water
398 137 360 120 240 0.44 (0.45) 1562 210 meth. water
BACK PRESSURE OF 221 kPa ON TEST A SPECIMEN FOR FOUR DAYS
355 221.06 403.4 261.1 142 0.82 (0.85) 1562 195 meth. water
508 353 561 400 162 0.89 (0.92) 1562 203 meth. water
450 295 368 226 142 0.84 (0.87) 1562 195 meth. water
TEST B
--- --- 0 -21 21 --- 255 130 air air
--- --- 0 -30 30 --- 268 138 air air
--- --- 0 -40 40 --- 278 147 air air
--- --- 0 -50 50 --- 288 154 air air
--- --- 0 -20 20 --- 243 124 air air
--- --- 47.6 0 47.6 --- 288 150 air air
TEST B SPECIMEN PURGED WITH CO2
--- --- 0 -50 50 --- 288 150 C0 2 air
30 3 40 7 32 0.45 (0.47) 1415 129 meth. air
54 30 103 54 50 0.47 (0.48) 1415 117 meth. air
106 66 155 91 64 0.52 (0.55) 1557 148 meth. air
157 95 211 124 87 0.54 (0.57) 1557 157 meth. air
219 145 198 135 63 0.49 (0.50) 1557 142 meth. air
197 134 174 120 54 0.60 (0.63) 1557 142 meth. air
160 116 93 73 20 0.63 (0.65) 1557 109 meth. air
253 140 283 153 130 0.43 (0.46) 1557 183 meth. water
BACK PRESSURE OF 154 kPa ON TEST B SPECIMEN FOR SIX DAYS
280 152 324 189 134 0.86 (0.98) 1557 188 meth. water
323 189 364 225 139 0.88 (0.99) 1415 188 meth. water
106 67 158 114 44 0.90 (0.96) 1557 136 meth. water
158 115 208 161 47 0.93 (0.99) 1730 142 meth. air
208 161 135 95 40 0.91 (0.98) 1557 135 meth. air
130 91 79 46 33 0.89 (0.95) 1557 127 meth. air
--- --- 77 4 73 --- 1416 157 meth. air
--- --- 73 -20 93 --- 1415 173 meth. air
--- --- 0 -20 20 --- 1557 117 meth. air
(1)
B-value in brackets is corrected for compliance of pore pressure measuring system
(2)
meth. = 1.75% methylcellulose in water (methocel)

140
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

Table 8-2 Soil and Fluid Properties used in Numerical Analyses

Description Symbol Value Units


Confining pressure (isotropic) ' m 100 kPa
Dry density d 1450 kg/m3
Shear modulus G 5.5x104 kPa
5
Bulk modulus of soil skeleton Kb 1.34x10 kPa
6
Fluid modulus (100% saturation) Kw 2.2 x10 kPa
-3
Permeability K 1.3 x10 cm/s
Porosity N 0.45 ---

141
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

FIGURES

Chapter 8
P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

142
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

0.8 (a)

Skempton 'B'
0.6

0.4

0.2

0
0.9 0.92 0.94 0.96 0.98 1
Degree of Saturation (Sr)
2000
(b)
P-wave velocity (m/s)

1500

1000

500

0
0 0.2 0.4 0.6 0.8 1
Skempton 'B'

Figure 8-1 Variation of (a) Skempton B with Degree of Saturation (S r ) and (b) P-wave velocity
(V p ) with Skempton B for Fraser River Sand, with D r =20% K b =92 MPa, G=56 MPa,
and porosity = 0.45 in accordance with Equations (1) and (2).

2.5
CSR(partial saturation)
CSR(fully saturation)

1.5

0.5
0 500 1000 1500 2000
P-wave velocity (Vp) (m/s)

Figure 8-2 Normalized cyclic strength ratio (CSR) of sand compared to P-wave velocity (V p ) as
measured in laboratory tests conducted with Toyoura sand. Developed using data from
Tsukamoto et al. (2002) and Ishihara et al. (1998). CSR is that which causes
liquefaction (double amplitude strain of 5%) in 20 cycles (for Relative Density of 40-
70%).

143
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

TRIAXIAL CELL FILLED


METAL PLATENS
WITH AIR OR WATER
SAMPLE

DIGITAL OSCILLOSCOPE
BENDER
ELEMENTS AMPLIFIER

RUBBER
MEMBRANE

Legend
VACUUM Regulator R
R
M TRIAXIAL Pressure
PT
CELL transducer

AIR AIR valve


1.5m

methocel M
R R

SAND
AIR

PT PT R
WATER

R C02

Figure 8-3 Laboratory test layout.

250
Shear Wave Velocity (m/s)

200

150

100

50
Dry Saturated
0
0 50 100 150 200
Mean Effective Stress (kPa)

Figure 8-4 Variation of shear wave velocity (V s ) with effective mean normal confining pressure
( m ) as measured in laboratory tests.

144
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

6
Input source motion at bottom of sample

P-wave arrival from shear wave


Amplitude

Measured wave at top of sample

(a)
-6
6
Input source motion at bottom of sample

P-wave arrival from shear


wave
Amplitude

Measured wave at top of sample

(b)
-6
0 0.5 1 1.5 2
Time (ms)

Figure 8-5 Typical P-wave time histories as measured in the laboratory. Traces (a) is for B = 0.48
with a calculated V p of 1415 m/s and trace (b) is for B = 0.95 with a calculated V p of
1557 m/s.

145
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

2000

Wave Velocity (m/s)


1500
Vp
1000

500 Dry
Vs
0
0 0.2 0.4 0.6 0.8 1
Skempton 'B'

Figure 8-6 Measured shear wave (V s ) and P-wave (V p ) versus Skempton B.

Figure 8-7 35 by 75 element mesh used for FLAC analyses.

146
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

air bubbles

(a) (b)

Figure 8-8 (a) Represents homogeneous-partial-saturation (HPS) where the air bubbles are small
and scattered throughout the void spaces. This is numerically modeled by reducing the
fluid modulus (K w ) at all the nodes in the model. (b) Represents non-homogeneous-
partial-saturation (NHPS) where there are only a few large bubbles at select locations.
This is numerically modeled by reducing the fluid modulus only at select nodes within
the model.

2500
FLAC - non homogeneous (NHPS)
FLAC - homogeneous (HPS)
Laboratory
Equation (2)
2000 Tamura Dr=65%
Ishihara Dr=60%
P-wave velocity (V p) (m/s)

1500

1000

500

0
0 0.2 0.4 0.6 0.8 1
Skempton B

Figure 8-9 Comparing V p versus Skempton B from both HPS and NHPS FLAC analyses to that
observed in laboratory tests and that calculated from Equation (2). Data from Ishihara
et al. (1998) and Tamura et al. (2002) that deviated from the correlation from Equation
(2) are also shown.

147
Chapter 8 P-Wave Measurements as an Indicator of Saturation in Centrifuge Tests

Figure 8-10 Typical variation of fluid modulus within NHPS FLAC grid. Small squares represent
locations with low fluid modulus representative of air bubbles, whereas all other
elements have a high fluid modulus of 2.2x109 Pa.

148
Chapter 9 Discussion and Conclusions

9 DISCUSSION AND CONCLUSIONS

9.1 Summary and Conclusions

Seismic design of major civil structures (bridges, dams and embankments) is moving
increasingly towards using performance design methodologies. A key aspect of performance
design is to determine the earthquake induced ground and structure movements. To do this,
designers are increasingly resorting to the use of dynamic numerical analyses. The research
presented in this dissertation involves development of these numerical design tools and
procedures for use in engineering practice for estimating the earthquake induced ground
deformations of soil that may potentially liquefy. The developed design methodology uses
effective stress and total stress soil constitutive models, UBCSAND and UBCHYST,
developed at the University of British Columbia (UBC). There are several assumptions and
approximations in these analysis procedures and due to this, a hybrid procedure that
combines the coupled effective stress analysis with a post-shaking static analysis check using
empirical soil strengths that have been back-calculated from past earthquake case histories
has been developed.

In Chapter 2, an overview of items contributing to liquefaction and potential flow


sliding are given. These are behaviours that are important for the proposed numerical model
to capture and important for the proposed users of the model to understand. It is noted that
sands have contractive tendencies when sheared at stress ratio less than the phase
transformation stress ratio, have dilative tendencies sheared at higher stress ratio, but always
contractive tendencies on unloading. The net response of cyclic shear loading of saturated
sands is nearly always a tendency for contraction. This in turn generates pore water
pressures, potential liquefaction of the sands (low effective stress and very low shear
moduli), and sets up a hydraulic gradient within the deposit. The gradients cause water flow
(generally upward or toward a free face) out of some zones and into others. Those zones
losing water generally contract and those gaining pore water expand. When there is a low
permeability barrier above the sand layer the inflow of pore water into the soil directly below
the barrier results in localization of both the expansion and shear strain. As the sand
expands, its potential shear-induced-dilative-tendencies decrease, and with sufficient

149
Chapter 9 Discussion and Conclusions

expansion, the critical state, at which shear strain will induce no further expansion, is
reached. This contraction and expansion can have a dramatic affect on both the effective
stress and the shear strength of the sand. This is particularly the case for loose sand under
low permeability barriers where expansion tends to be localized and strengths can go to very
low values, or at the limit, zero when a water interlayer forms. Most of the concepts in
Chapter 2 have been discussed by others in some manner in the past. However, the
framework and emphasis of the discussion is important and is that of the author.

Soil mixing, another mechanism that can lead to significant strength reduction of
layered liquefiable soils and liquefied soils, is investigated in Chapter 3. Many natural
cohesionless soils and man-made hydraulically placed soils are layered. When the fine and
coarse layers mix, the resulting residual shear strength is much lower than that of the
individual layers at the same void ratio. This mixing provides explanation for the 1970 flow
liquefaction failure observed at the Mufulira Mine site and, at least partially, accounts for the
low end-of-failure residual shear strength back-calculated for the Lower San Fernando Dam
failure. Mixing is not a new concept, but is a mechanism that can both result in liquefaction
and lower the undrained strength of liquefied soils.

The author proposed the concept of cyclic mixing and developed rudimentary cyclic
shear tests to investigate the concept. The rudimentary cyclic shear tests were carried out
within a 12mm internal diameter vinyl tube. During the large strain cyclic shearing the
movement of the soil particles was clearly visible. The ratio of the grain-size of the coarse
and fine layers was shown to be a key factor in a layered soils susceptibility to mixing. A
preliminary postulation by the author is that cohesionless soils may be susceptible to cyclic
mixing when ((D 15 ) coarse layer / (D 85 ) fine layer ) > 3 to 4. It is also postulated that mixing will
more readily occur when effective stresses and related inter-particle contact forces are low.
Mixing is probably more common as a post-liquefaction mechanism and helps explain the
low post-liquefaction strengths and large run-out distances observed in some instances.
Within a post-liquefaction low effective stress low strength environment both mechanical
shear and turbulent flow may induce mixing. The occurrence of mixing induced by turbulent
flow is a concept postulated by the author following examination of the shake table test
videos by Kokusho (http//:www.civil.chuo-u.ac.jp/lab/doshitu/index.html).

150
Chapter 9 Discussion and Conclusions

The post-liquefaction residual strengths back-calculated from case histories are


discussed in Chapter 4. These are important because they are used in current engineering
practice and also because they provide the input parameters for the proposed hybrid
numerical procedure that is the topic of this thesis. Experience has shown that strengths
back-calculated from laboratory tests tend to give strengths that are too high. This is
postulated to be due to the laboratory testing process not accounting for void redistribution
and mixing effects. If the in-situ soil is truly believed to behave in an undrained manner then
the strengths from laboratory tests should be representative. In the field, the soil is generally
not truly undrained and the residual strength is dependent on many variables (soil
density, soil fabric, stratigraphy, initial stresses, duration of shaking, permeability /
groundwater flow regime, method of analyses used, etc.) and is as much a site property as it
is a soil property. Correlating the residual strength or strength ratio with soil density or
penetration resistance is a gross simplification and consequently there is much scatter and
uncertainty in the data.

Chapter 5 introduces the constitutive models UBCSAND and UBCHYST that are
key parts (especially UBCSAND) of the proposed hybrid analysis procedure. UBCSAND,
when run within the program FLAC is a coupled effective stress model that considers shear
induced soil skeleton volume changes, their effect on pore water pressure, and the resultant
pore water flow from one element to another. As indicated in Chapter 2, these are key
behaviours needed to capture soil liquefaction and related soil strength changes. The
UBCSAND program was developed by Dr. Byrne and his students. New features added to
UBCSAND as part of this research include:

Elimination of numerical instability for certain cases with static bias;

Improvement in the simulation of cases with large static bias;

Improvement in the liquefaction behaviour by adding a dilation delay (period


during which shear does not induce volume change) thus allowing large shear
strains to develop when the normal effective stress and shear stress is near
zero, in agreement with behaviour observed in laboratory test data (Figure
5-10),

151
Chapter 9 Discussion and Conclusions

Introduction of a dilation cut-off which set shear induced dilation to zero


when an element reached a pre-described critical volume or void ratio
(analogous to the dilation angle going to zero when the soil reaches the critical
state); and,

Improvements in the modelling of dense soils by making the phase


transformation stress ratio a function of soil density and changing the
correlation between peak friction angle and soil density.

UBCHYST is a simple robust total stress constitutive model often used together
with UBCSAND for the numerical modelling of soil structures subjected to earthquake
shaking. UBCSAND is used for portions of the soil profile deemed susceptible to
liquefaction and UBCHYST is used for those deemed not susceptible. The advantage of the
UBCHYST model over a simpler Mohr Coulomb model is the non-linear hysteretic loops
developed by varying the tangent shear modulus during loading and unloading. This more
closely replicates the behaviour of real soil and reduces the Rayleigh damping requirements;
thus increasing analysis speed. The initial algorithm for UBCHYST was developed by
Professor Byrne while the author coded UBCHYST as a constitutive model in FLAC, made
improvements regarding unload/reload loops, developed options for strength reduction as a
function of strain or number of cycles, and developed documentation explaining the model.

The proposed hybrid effective stress total stress design approach is presented in
Chapter 6. Dynamic numerical analyses using the UBCSAND constitutive model may be the
state-of-the-art in predicting liquefaction induced deformations. However, there are still
shortcomings, including: element size / localization effects, underestimation of post-
liquefaction consolidation, no out-of-plane (3D) shaking, etc. In addition to the modelling
shortcomings, there is also the difficulty of characterizing the soil stratigraphy with accuracy
at any given real site. Case history derived post-liquefaction soil strengths also have large
uncertainties associated with them and they are used in current engineering practice largely
because there are no alternatives. When considering all this, it is proposed that a good design
should consider both approaches and the hybrid effective stress total stress design approach
described in this chapter is the result. A description of the procedure and step-by-step details
on implementing it are given in Chapter 6.

152
Chapter 9 Discussion and Conclusions

Chapter 7 provides examples of both the calibration and validation of the


UBCSAND model and the proposed hybrid analysis procedure.

Laboratory tests are one of the few cases where the loading, stress and pore pressure
histories can be controlled and are known. Section 7.2 shows comparisons between a single
element with the UBCSAND (version UBCSAND1v02) model and cyclic simple shear tests
by Sriskandakumar (2004). It is shown that the new version of UBCSAND with the
proposed revisions gives a reasonable match of both loose and dense sand samples when
there is no static bias. With a static bias the match is not as good; however, the behaviour
mechanisms are matched.

Section 7.3 describes a calibration of the model to empirical liquefaction triggering


charts. The empirical liquefaction triggering charts have been developed from numerous
case histories and their use is accepted current design practice. This section shows that
UBCSAND can be calibrated to simulate the liquefaction triggering behaviour inferred by
these field case history derived correlations. Calibrating to these charts is commonly done
when using UBCSAND for commercial design projects.

Section 7.4 describes some early, 2004, one dimensional (infinite slope) column
tests that the author carried out to demonstrate that UBCSAND/FLAC could simulate the
post-shaking flow failure phenomena. To the authors knowledge this is the first numerical
simulations of this type of behaviour. This simulation demonstrates: the ability of
UBCSAND/FLAC to generate pore pressure gradients and flow, the mechanism of some
elements losing pore water and decreasing in volume and some elements gaining pore water
and increasing in volume, and the mechanism of volume increase and eventual strength loss
that occurs below a low permeability barrier layer that overlies liquefiable sands.

Section 7.5 describes some work carried out for the seismic retrofit design of the
George Massey tunnel. This is one of the earlier commercial uses of UBCSAND and
included a centrifuge test program for validation. The author supervised and was responsible
for the analyses. The numerical simulations of the centrifuge tests were Class A
predictions (carried out before the centrifuge tests were carried out) and gave good
agreement with the actual test performance.

153
Chapter 9 Discussion and Conclusions

Section 7.6 described some numerical modeling of centrifuge tests carried out by
C-CORE in Newfoundland as part of the UBC Liquefaction Initiative. These are some of the
earlier centrifuge tests to incorporate an embankment slope with a low permeability soil
barrier overlying loose liquefiable sand, and possibly the first centrifuge test to demonstrate
flow failure after end-of-shaking. This is a behaviour that has been observed in several field
case histories. The author showed that the UBCSAND/FLAC numerical model could
simulate the observed post-shaking failure mechanism when shear-induced dilation was
curtailed upon the element volume exceeding a preset threshold. This is analogous to shear
induced volume change being zero when the soil reaches the critical state. These centrifuge
tests were suspected of not being fully saturated. This initiated the triaxial testing program
and numerical analyses described in Chapter 8 of this dissertation.

The shake table tests by Kokusho (2003) were important because they illustrated the
void redistribution mechanism that explains the post-shaking failures so often observed in
field case histories, and because they rekindled interest by the research and design
community in the void redistribution mechanism. Section 7.7 describes numerical modelling
by the author that demonstrated that the observed deformation and post-shaking mechanism
could be simulated using the UBCSAND constitutive model within FLAC.

The 1971 failure of the Lower San Fernando Dam in California is one of the best
documented liquefaction-induced-failure case histories available. A key aspect of this case
history is that it failed approximately 30 seconds after the end-of-shaking. Therefore, the
failure was caused by delayed soil strength reduction and not by earthquake inertial forces.
Many attempts by others to numerically model this failure have been able to approximately
simulate the deformation pattern but, to the authors knowledge, none have successfully
simulated the post-shaking shear failure mechanism. Section 7.8 describes the authors work
on numerical modeling of the Lower San Fernando Dam using the proposed hybrid
procedure. It is shown that the post-failure mechanism can be simulated by the UBCSAND
coupled effective stress model. It is also shown that carrying out the proposed post-failure
static analysis check, using the Idriss and Boulanger, 2008 post-liquefaction shear strength
correlations, within the zones that liquefied during the strong shaking portion of the effective
stress analysis, also results in flow failure.

154
Chapter 9 Discussion and Conclusions

The author was able to simulate post-shaking flow failure in the analyses described
in Sections 7.4 (infinite slope column with low-permeability barrier), 7.6 (C-Core centrifuge
tests), 7.7 (Kokusho shake table simulation), and 7.8 (Lower San Fernando Dam failure). In
each of these cases, when the same analyses are repeated with flow not permitted (undrained
condition) or repeated without the low permeability layers, then post-shaking failure did not
occur. This highlights the importance of being able to simulating pore water flow and void
redistribution mechanisms when numerically modelling potentially liquefiable earth
structures subjected to earthquake shaking.

Chapter 8 is on a slightly different theme then the rest of the thesis. However, the
observations and conclusions derived are believed to be of technical interest and, therefore,
were included in the main body of the dissertation. Centrifuge tests were carried out as part
of the UBC Liquefaction Research Initiative (NSERC Grant 246394) for the purpose of
validating the UBCSAND numerical model. During this work it was noted that the
behaviour in the centrifuge did not match that predicted by the numerical models and it was
suspected that the pore fluids in the centrifuge were not fully saturated, even though P-wave
velocity (V p ) measurements were indicative of full saturation. To check why this may be the
case, triaxial tests, that simulated the centrifuge saturation procedure, were carried out. The
triaxial cells had bender elements in the top and bottom platens and thus shear wave velocity
(V s ) and P-wave velocity (V p ), and Skempton B could be measured during the testing. The
results from the testing showed that when the measured Skempton B in the triaxial sample
was high (representative of full saturation), the corresponding V p was also high. However, a
high V p did not always correspond to a high Skempton B. It is postulated that the behaviour,
in the later case, is due to the air bubbles which give the low Skempton B, being widely
spaced, thus, allowing the P-waves to find travel paths through the medium that bypass the
bubbles (thus giving the high velocity). From this, two types of partial saturation have been
defined: homogeneous-partial-saturation (HPS) whereby the air bubbles are numerous, very
small and included within each void space throughout the specimen, and non-homogeneous-
partial-saturation (NHPS) whereby the air bubbles are much larger and are only present
within some pore spaces. For HPS, the V p will be a function of the Skempton B and average
saturation (S r ) av , whereas, for NHPS, V p is independent of Skempton B and average
saturation. With NHPS, high V p values indicative of high saturation are possible, even

155
Chapter 9 Discussion and Conclusions

though the B value may be relatively low. NHPS conditions are postulated to be more
prevalent in centrifuge tests where viscous solutions, such as methocel are often used in lieu
of water. In addition to the laboratory test, the behaviour was also numerically simulated
using the program FLAC. This later simulation provided confidence in the postulated
explanation for the behaviour.

When the triaxial specimen was placed under backpressure for several days this
resulted in air going into solution and increased saturation and related Skempton B values.
This procedure is standard practice in the saturation of laboratory specimens, and it is
proposed that the procedure also be considered for saturating centrifuge test specimens.

9.2 Recommendations

9.2.1 Recommendations based on the Work carried out for this Dissertation

(1) The design process outlined in this dissertation is a viable design tool for use in
engineering practice and its use is recommended. By carrying out the numerical
modelling significant insight into the design problems can be obtained. In addition,
the calculated displacements are likely more reliable than those obtained using
simplified methods based on Newmark-type procedures (Newmark 1965) or those
obtained using empirical correlations such as those by Youd et al. 2002).

(2) For design work, the model can be calibrated to published liquefaction triggering
and K correlations (charts); however, the post-liquefaction triggering response, and
response to a static shear bias, is important and is not adequately captured by these
charts. For these aspects, the use of simple shear laboratory tests is recommended.
It is suggested that tests with similar stress paths, to what will be experienced in the
design problem, be carried out.

(3) On important large projects, consideration should be given to carrying out physical
model tests to calibrate and validate the numerical model. Centrifuge testing and
large shake table tests may be useful for this purpose.

(4) When carrying out centrifuge tests, measured P-wave velocities may not correlate
well with saturations. Placing a specimen under back pressure for several days
results in air going into solution, which increases saturation and related B values.

156
Chapter 9 Discussion and Conclusions

This procedure is standard practice in the saturation of laboratory specimens, and it


is recommended that the procedure also be considered for saturating centrifuge test
specimens.

9.2.2 Recommendations for Future Research

UBCSAND

(1) Further work is required for improving the model in conditions with large static
bias. Much of the deformation during earthquake shaking is related to marching
due to the static bias. More laboratory tests with stress paths that simulate that
which occurs in the field are required to assist with model development.

(2) In many instances, the volumetric strain due to post-liquefaction consolidation may
be much larger than currently calculated by UBCSAND, especially in level ground
conditions. This could be important as with the correct post-liquefaction volumetric
strain, more water would be available for re-distributing, and there would be more
potential for water films, etc. A version of UBCSAND with improved post-
liquefaction consolidation behaviour should be developed. Introducing a volumetric
cap or yield envelope may be required to do this. An option in FLAC to have a
rigid (infinitely high) fluid modulus may also help.

(3) In this dissertation, analyses were carried out that showed that the UBCSAND
constitutive model could capture the post-shaking flow failure observed in the 1971
flow failure of the Lower San Fernando Dam. It is suggested that analyses using the
same methodology should be carried out for the Upper San Fernando Dam where
flow failure did not occur during the same earthquake with similar soil conditions.

(4) The model is a compromise - a balance between simplicity, practicality, robustness


and ability to capture field behaviour. This should be considered when
implementing changes.

Mixing

The importance of mixing on post-liquefaction flow failure run-out distances and


residual strength is believed to be underestimated. Further study of the mixing phenomena
should be undertaken. This could include more detailed assessment of which grain-size
157
Chapter 9 Discussion and Conclusions

variations, shear strain values, number of cycles, etc. would induce mixing and which would
not. Also, it is observed that much mixing occurs between layers of non-plastic materials
when there is liquefaction and extensive pore water void redistribution perhaps
development of special laboratory apparatus (or shaking table tests) are needed to adequately
study mixing.

Use of Large Shaking Tables

Large-sized shaking tables provide opportunity to study pore water void


redistribution phenomena and the mixing between layers. Tests similar to those by Kokusho
(Kokusho 1999, 2000, 2003; and Kokusho and Kabasawa 2003) could be repeated on a much
larger real-life scale with detailed measurements (using video recording or photographs, if
clear sides are used) of void changes etc.

Extension of Analyses Tools into three dimensions

Most real soil-structure interaction design problems (bridge piers, building


foundations, etc.) have a strong three dimensional component and the computing power
necessary for these analyses is currently available. Making the UBCSAND/UBCHYST type
design tools three dimensional, is a logical goal.

Residual or Post-liquefaction Strength of Soils for Design

Current empirical case-history based procedures are inaccurate and improved


correlations and methodologies are needed. Variables other than those in the current
correlations may be important. Static bias and stratigraphy are expected to have a significant
effect. The use of coupled effective stress numerical analyses to back-analyze case histories
and generic design sections might prove helpful in developing new design correlations or
methodologies.

Use of particulate models (such as, ITASCAs PFC3D program)

Current research using PFC3D to simulate laboratory tests suggests that the use of
particle models can capture the behaviour observed in laboratory triaxial and simple shear
tests (Pinheiro et al. 2008; Dabeet 2010) and provide useful assistance in understanding soil
behaviour and liquefaction phenomena. Improvement in particulate codes and in computing
power are areas where further development is required. Using available codes to further

158
Chapter 9 Discussion and Conclusions

develop understanding in soil behaviour and to develop engineering design tools is another
suggested area of research.

Liquefaction Triggering

An improved understanding of the effect of static bias on triggering is required.


Current practice is to set K (a factor to allow for static bias) to unity which implies that
there is no effect. However, both laboratory testing and field experience indicate that there
can be a quantifiable effect. A simple procedure may not be possible and numerical models
may be required to capture the behaviour, or else numerical models may be able to assist in
developing an improved simplified design procedure.

159
References

REFERENCES

Adalier K., Elgamal A., Meneses J., and Baez J.I. 2003. Stone column as liquefaction
countermeasure in non-plastic silty soils. Soil Dynamics and Earthquake
Engineering, 23(7): 571-584.
Amini, F., and Qi, G.Z., 2000. Liquefaction testing of stratified silty sands. J. of Geotech.
Geoenviron. Eng., ASCE, Vol. 126, No. 3, March, pp. 208-217.
Atigh, E. and Byrne, P.M. 2004. Liquefaction flow of submarine slopes under partially
undrained conditions: an effective stress approach. Canadian Geotech. J., V. 41, pp.
154-165.
Barrett, J.R., Byrne, P.M, 1988. Review of mining below TD#3, Mufulira Mine, 1988.
Report to Mines Safety Department, Republic of Zambia and Zambia Consolidated
Copper Mines Ltd.
Baziar, M.H., and Dobry, R., 1995. Residual strength and large deformation potential of
loose silty sands. J. Geotech. Engineering Div., ASCE, Vol. 121 (12), pp. 896-906.
Beaty, M.H. 2001. A synthesized approach for estimating liquefaction-induced
displacements of geotechnical structures. Ph.D. Thesis, University of British
Columbia, Dept. of Civil Eng.
Beaty, M. 2009. Summary of UBCSAND constitutive model: Versions 904a and 904aR.
Draft Document. Beaty Engineering LLC.
Beaty, M., and Byrne, P.M. 1998. An effective stress model for predicting liquefaction
behaviour of sand. Geotechnical Earthquake Engineering and Soil Dynamics III. P.
Dakoulas, M. Yegian, and R Holtz (eds.), ASCE, Geotechnical Special Publication
75 (1), pp. 766-777.
Beaty, M. H. and Byrne, P. M. 1999. A synthesized approach for modeling liquefaction and
displacements. In Proc., Conf., FLAC and Numerical Modeling in Geomechanics,
Edited by Detourneayand Hart, R., A. A. Balkema, Rotterdam, pp. 339-347.
Beaty, M.H. and Byrne, P.M. 2011. UBCSAND constitutive model Version 904aR.
Document report: UBCSAND Constitutive Model on Itasca UDM Web site:
http://www.itasca-udm.com/pages/continuum.html
Boulanger, R. W., and Truman, S. P. 1996. Void redistribution in sand under post-
earthquake loading. Can. Geotech. J., 33, 829-834.
Bray, J.D. and Travasarou, T. 2007. Simplified procedure for estimating earthquake-induced
deviatoric slope displacements. J. of Geotech. and Geoenv. Engineering, ASCE,
April, Vol. 133, No. 4, pp. 381-392.
Byrne, P.M., 1989. Liquefaction review report, Mufulira mine, Zambia. Report to Mines
Safety Department, Republic of Zambia and Zambia Consolidated Copper Mines
Ltd.

160
References

Byrne, P.M., 1991. Evaluation and Use of Residual Strength in Seismic Safety Analysis of
Embankments. Earthquake Spectra, Vol. 7, No. 1, Feb., pp. 145-148.
Byrne, P.M., and Beaty, M. 1997. Post-liquefaction shear strength of granular soils:
Theoretical/conceptual issues [Keynote Paper]. In Proceedings of the Workshop on
Shear Strength of Liquefied Soils, Urbana, NSF Grant CMS-95-31678, pp 9-39.
Byrne, P.M., Park, S.S., Beaty, M., Sharp, M., Gonzalex, L., Abdoun, T. 2004. Numerical
modeling of dynamic centrifuge tests. 13th World Conf. on Earthquake Eng.,
Vancouver, B.C., paper 3387.
Byrne, P.M., Naesgaard, E., and Seid-Karbasi, M. 2006. Analysis and design of earth
structures to resist seismic soil liquefaction. 59th Canadian Geotechnical Conf.,
Vancouver.
Byrne, P.M., Naesgaard, E., and Beaty, M.H., 2008. State of art dynamic liquefaction
analysis procedures. Presentation to ITASCA FLAC conference, Minnesota, Aug.
Casagrande, A. 1936. Characteristics of cohesionless soils affecting the stability of slopes
and earth fills. J. Boston Society of Civil Engineers, pp. 257-276.
Castro, G. 1975. Liquefaction and cyclic mobility of saturated sands. J. Geotech. Eng. Div.,
ASCE, Vol. 101, No. GT6, Proc. Paper 11388, 551-569.
Castro, G. 1995. Empirical Methods in Liquefaction Evaluation. Primer Ciclo de
Conferencias Internationales Leonardo Zeevaert, Mexico City. Castro, G., Keller,
T.O., and Boynton, S.S. 1989. Re-evaluation of the Lower San Fernando Dam, Rpt.
1 Vol. 1 and 2. US Army Engineer Waterways Experiment Station, Contract Rpt.
GL-89-2.
Cetin, K.O., Seed, R.B., Kiureghian, A.D., Tokimatusu, K., Harder, L.F. Jr., Kayen, R.E.,
and Moss, R.E.S. 2004. Standard Penetration Test-Based Probabilistic and
Deterministic Assessment of Seismic Soil Liquefaction Potential. J. of
Geotechnical and Geoenvironmental Engineering, ASCE, 130(12): 1314-1340.
Chaney, R. 1978. Saturation effects on the cyclic strength of sands. Earthquake Engineering
and Soil Dynamics, 1, Proc., ASCE Geotechnical Engineering Div. Specialty
Conference, Pasadena, CA, 342-359.
Cundal, P. and Board, M. 1988. A microcomputer program for modeling large-strain
plasticity problems. Proc., 6th Int. Conf. Numerical Methods in Geomechanics,
Balkema, Rotterdam, 2101-2108.
Dabeet, A., Wijewickreme, D. and Byrne, P. 2010. Evaluation of the stress-strain
uniformities in the direct simple shear device using 3D discrete element modeling.
In proceeding of Canadian Geotechnical Conf., GeoCalgary, Calgary, Alberta,
Canada, September.
Dismuke, J. N. 2003. Aspects of the cyclic loading behavior of saturated soils. M.S. thesis,
Univ. of California, Davis, Calif.
Dashti, S. 2009. Toward developing an engineering procedure for evaluating building
performance on softened ground. Ph.D. Thesis, University California, Berkeley.

161
References

Elgamal, A., Yang, Z., Parra, E., and Dobry, R. 1999. Modeling of liquefaction-induced
shear deformations. Second International Conference on Earthquake Geotechnical
Engineering, Lisbon, Portugal, 21-25 June, Balkema Publisher.
Finn, W.D.L., Lee, K.W., Martin, G.R. 1977 An effective stress model for liquefaction. J.
Geotech. Eng. Div., ASCE 103 (GT6), pp. 517-533.
Finn, W.D.L., Bhatia, S.K., and Pickering, D.J. 1982. The cyclic simple shear test. Soil
Mechanics Transient and Cyclic Loads. Edited by G.N. Pande and O.C.
Zienkiewicz, pp. 583-605.
Finn, W.D.L., Yogendrakumar, M., Yoshida, N., Yoshida H. 1986. TARA-3: A program for
nonlinear static and dynamic effective stress analysis. Soil Dynamics Group,
University of B.C., Vancouver, B.C.
GDS Instruments Ltd. 2002. GDS BES - GDS Bender element system software handbook
and GDS Bender element hardware handbook, GDS Instruments Ltd., Hampshire,
UK.
Hamada, M. 1992. Large ground deformations and their effects on lifelines: 1964 Niigata
Earthquake. Proc., Lifeline Performance during Past Earthquakes, V. 1: Japanese
Case Studies Tech. Rep NCEER-92-0001, Buffalo, N.Y., 3/1-3/123.
Horsfield, D.W., and Been, K., 1989. Field and laboratory testing of Tailings Dump #3,
Residual Inrush and Sandfill Tailings. Mufulira Copper Mine, Golder Associates
report to Zambia Consolidated Copper Mines Ltd.
Idriss, I.M., and Boulanger, R.W. 2006. Semi-empirical procedures for evaluating
liquefaction potential during earthquakes. Soil Dynamics and Earthquake
Engineering 26, 115-130.
Idriss, I.M., and Boulanger, R.W. 2007. SPT- and CPT-based relationships for the residual
shear strength of liquefied soils. Proc. 4th Int. Conf. on Earthquake Geotechnical
Engineering, Thessaloniki, Greece, June.
Idriss, I.M., and Boulanger, R.W. 2008. Soil liquefaction during earthquakes. EERI MNO-
12.
Ishihara, K. 1984. Post-earthquake failure of a tailings dam due to liquefaction of the pond
deposit. Proc. Inter. Conf., Case Histories in Geotechnical Eng., Rolla, Missouri, V.
3, pp. 1129-1143.
Ishihara, K. 1993. Liquefaction and flow failure during earthquakes. J. Geotechnique, V.
43(3), pp. 351-415.
Ishihara, K. and Yoshimine, M. 1992. Evaluation of settlements in sand deposits following
liquefaction during earthquakes. Soils and Foundations. Vol. 32(1): 173-188.
Ishihara, K., Y. Huang and H. Tsuchiya 1998. Liquefaction resistance of nearly saturated
sand as correlated with longitudinal wave velocity. Poromechanics: A Tribute to
Maurice A. Biot, Balkema, 583 586.
ITASCA, 2002. FLAC, Fast Lagrangian Analysis of Continua, A Users Manual, Version
4.0. ITASCA Consulting Group, Inc., Minneapolis, Minnesota.

162
References

ITASCA, 2003. PFC3D (Particle Flow Code in 3 Dimensions), Version 3.0. ITASCA
Consulting Group, Inc., Minneapolis, Minnesota.
ITASCA, 2005. FLAC - Fast langrangian analysis of continua, Version 5.0. ITASCA
Consulting Group Inc., Minneapolis, Minnesota.
ITASCA, 2008. FLAC - Fast langrangian analysis of continua, Version 6.0 ITASCA
Consulting Group Inc., Minneapolis, Minnesota.
ITASCA, 2008. FLAC3D - Fast langrangian analysis of continua in 3 dimensions, Version
4.0 ITASCA Consulting Group Inc., Minneapolis, Minnesota.
Khoei, A.R., Azami, A.R., Haeri, S.M. 2004. Implementation of plasticity based models in
dynamic analyses of earth and rockfill dams; A comparison of Pator-Zienkiewicz
and cap models. Comp0uters and Geotechnics, 31, 385-410.
Khoei, A.R., Anahid, M., Zarinfar, M., Ashouri, M., and Pak, A. 2011. A large plasticity
deformation of unsaturated soil for 3D dynamic analyseis of Lower San-Fernando
Dam. Asian J. of Civil Eng. (Buildings and Housing) Vol. 12 No. 1, 1-25.
Kokusho, T., 1999. Water film in liquefied sand and its effect on lateral spread. J. Geotech.
Geoenviron. Eng. Vol. 125 No. 10, 817826.
Kokusho, T. 2000. Mechanism for water film generation and lateral flow in liquefied sand
layer. Soils & Foundations, 40(5), 99-111.
Kokusho, T. 2002. Correlation of pore-pressure B-value with P-wave velocity and Poissons
ratio for imperfectly saturated sand or gravel. Soils and Foundations, 40(4), 95-102.
Kokusho, T. 2003. Current state of research on flow failure considering void redistribution
in Liquefied deposits. Soil Dynamics and Earthquake Engineering 23, 585-603.
Kokusho, T., Yoshikawa, T., Suzuki, K., and Kishimoto, T., 2003a. Post-liquefaction shear
mechanism in layered sand by torsional shear tests. 12th Panamerican Conf. on Soil
Mech. and Geotech. Eng., June, pp. 1045-1050.
Kokusho, T, and K. Kabasawa, K. 2003b. Energy approach to flow failure and its
application to flow due to water film in liquefied deposits. Proc. of International
Conference on Fast Slope Movements, Prediction and Prevention for Risk
Mitigation, Naples, 297-302, May 2003.
Kulasingam, R. 2003. Effects of Void Redistribution on Liquefaction- Induced
Deformations. Ph.D. thesis, Univ. Calif., Davis, 460 pp.
Kulasingam, R., Malvick, E.J., Boulanger, R.W., and Kutter, B.L., 2004. Strength loss and
localization of silt interlayers in slopes of liquefied sand. J. of Geotech. Geoenviron.
Eng., ASCE, Vol. 130, No. 11, November 1, pp. 1192-1202.
Lysmer, J., Udaka, T., Tsai, C.F. and Seed, H.B. 1975. FLUSH - A computer program for
approximate 3-D analysis of soil-structure interaction problems. EERC, University
of California, Berkeley, CA. Report NO. EERC 75-30.
Malvick, E. J., 2005. Void Redistribution-Induced Shear Localization and Deformation in
Slope. Ph.D. Thesis, University of California, Davis.

163
References

Matasovic, N., and Gustavo G., 2007. D-MOD2000 A computer program for seismic
response analyses of horizontally layered soil deposits, earthfill dams and solid
waste Landfills, GeoMotions, LLC., Washington.
Naesgaard, E., Yang, D., Byrne, P.M. and Gohl, B. 2004. Numerical analyses for the
seismic safety retrofit design of the immersed-tube George Massey Tunnel. Proc.
13th World Conf. on Earthquake Engineering, Vancouver, paper 112, August.
Naesgaard, E., and Byrne, P. M., 2005. Flow Liquefaction due to Mixing of Layered
Deposits. Proc. of Geot. Earthquake Eng. Satellite Conf., TC4 Committee,
ISSMGE, Osaka, Japan, Sept.
Naesgaard, E., Byrne, P. M., Seid-Karbasi, M., and Park, S. S., 2005. Modeling flow
Liquefaction, its mitigation, and comparison with centrifuge tests. Proc. of Geot.
Earthquake Eng. Satellite Conf., TC4 Committee, ISSMGE, Osaka, Japan, Sept.
Naesgaard, E., Byrne, P.M., and Seid-Karbasi, M. 2006. Modeling flow liquefaction and
pore water redistribution mechanisms. 8th NCEE, San Francisco, April.
Naesgaard, E. and Byrne, P.M., 2007. Flow liquefaction simulation using a combined
effective stress - total stress model. 60th Canadian Geotechnical Conference,
Canadian Geotechnical Society, Ottawa, Ontario, October.
Naesgaard, E., Byrne, P.M., and Wijewickreme, D., 2007. Is P-Wave Velocity an Indicator
of Saturation in Sand with Viscous Pore Fluid?. International Journal of
Geomechanics, ASCE. Vol. 7 No. 6 Nov/Dec. pp 437-443.
Naesgaard, E., Beaty, M.H., and Byrne, P.M., 2009. Performance-based design of
potentially liquefiable embankments using a combined effective stress total stress
model. IS-2009 Tokyo, Performance-Based Design in Earthquake Geotechnical
Engineering Kokusho, Tsukamoto, and Yoshimine (eds) 2009 Taylor & Francis
group London, ISBN 978-0-415-55614-9.
National Research Council (NRC) 1985. Liquefaction of soils during earthquakes. NCR
report CETS-EE-001, National Academy Press, Washington, D.C.
NCEER 1997. Proceedings of the NCEER workshop on Evaluation of Liquefaction
Resistance of Soils. Edited by Youd, T. L., Idriss, I. M., Technical Report No.
NCEER-97-0022, December 31, 1997.
Negussey, D., Wijewickreme, W.K.D. and Vaid, Y.P. 1988. Constant-volume friction angle
of granular soils. Can. Geotech. V. 25 (1), pp. 50-55.
Newmark, N. 1965. Effects of earthquakes on dams and embankments. Geotechnique Vol
15, No. 2, pp 139-160.
Olson, S. M. and Stark, T. D., 2002. Liquefied strength ratio from liquefaction flow failure
case histories. Canadian Geot. J., Vol 39, pp. 629647.
Park, S.S. 2005. A two mobilized-plane model and its application for soil liquefaction
analyses. Ph.D. thesis, University of British Columbia.

164
References

Parra, E. 1996. Numerical modeling of liquefaction and lateral ground deformation


including cyclic mobility and dilation response in soil systems. Ph.D. Thesis, Dept.
of Civil Engineering, RPI, Troy, NY.
Prevost, J.H. 2002. Dynaflow A nonlinear transient finite element analysis program,
Version 2002. Release 01.A., Dept. of Civil Engrg. & Operation Research,
Princeton Universtiy, Princeton, NJ. First Release, 1981.
Phillips, R.,, Tu, M., and Coulter, S., 2004. C-CORE. Earthquake Induced Damage
Mitigation from Soil Liquefaction. Data report Centrifuge Test CT5. For
University of British Columbia. C-CORE Report R-04-068-145, December.
Phillips, R., and Coulter, S., 2005. COSTA-C Centrifuge Test Data Report, C-CORE
Report R-04-082-075, January.
Pinheiro, M., Wan, R.G., Li, Q. 2008. Drained-undrained response and other fundamental
aspects of granular materials using DEM. In proceeding of Canadian Geotechnical
Conference,GeoEdmonton, Edmonton, Alberta, Canada, September.
Poulos, S.J., Castro, G., and France, W., 1985. Liquefaction evaluation procedure. J.
Geotechnical Engineering Div., ASCE, Vol. 111, No.6, pp. 772-792.
Roscoe, K.H. 1970. The tenth Rankine Lecture: The influence of strains in soil mechanics. J.
Geotechnique, V. 20, pp.129-170.
Rowe, P.W. 1962. The stress-dilatancy relation for static equilibrium of an assembly of
particles in contact. Proc. R. Soc. London Ser. A 269, pp. 500-527.
Seed, H.B., 1987. Design problems in soil liquefaction, J. Geotechnical Eng. Div., ASCE,
Vol. 113, No.8, 827-845.
Seed, H.B., Lee, K.L., Idriss, I.M., and Makdisi, F., 1973. Analysis of the Slides in the San
Fernando Dams during the Earthquake of Feb. 9, 1971. Earthquake Engineering
Research Center, Rpt. EERC 73-2, June.
Seed, H.B., Tokimatsu, K., Harder, L.F., Chung, R.M., 1984. The influence of SPT
procedures in soil liquefaction resistance evaluations. Report No. UCB/EERC-
84/15. Earthquake Engineering Research Center, Berkeley, California, 1984.
Seed, H.B., Seed, R.B., Harder, L.F., and Jong, H.L., 1989. Re-evaluation of the Lower San
Fernando Dam, Rpt. 2. US Army Engineer Waterways Experiment Station, Contract
Rpt. GL-89-2.
Seed, R.B., and Harder, L.F., Jr., 1990. SPT-based analysis of cyclic pore pressure
generation and undrained residual strength., In Proc. of the H.B. Seed Memorial
Symposium, Bi-Tech Publishing Ltd., Vol. 2, pp. 351376.
Seid-Karbasi, M., 2008. Effects of void redistribution on liquefaction-induced ground
deformations in earthquakes: a numerical investigation. Ph.D. thesis, University of
British Columbia.
Seid-Karbasi, M., and Byrne, P.M., 2004. Liquefaction, lateral spreading and flow slides.
Proc. 57th Canadian Geotechnical Conf., Session 2c, pp. 23-30.

165
References

Seid-Karbasi, M. Byrne, P.M., Naesgaard, E., Park, S.S., Wijewickreme, D., and Phillips, R.,
2005. Response of Sloping Ground with Liquefiable Materials during an
Earthquake: a class A prediction, 11th IACMAC Conf., Italy.
Sento, N., Kazama, M., Uzuoka, R., Ohmura, and Ishimaru, M., 2004. Possibility of
postliquefaction flow failure due to seepage. J. of Geotech. Geoenviron. Eng.,
ASCE, Vol. 130, No. 7, pp. 707-716.
Skempton, A.W. 1954. The pore pressure coefficient A and B. Geotechnique, 4(4), 143-147.
Sriskandakumar, S., 2004. Cyclic loading response of Fraser River Sand for validation of
numerical models simulating centrifuge tests. M.A.Sc. Thesis, University of British
Columbia, Vancouver, B.C., 159 p.
Stark, T.N. and Mesri, G. 1992. Undrained shear strength of liquefied sands for stability
analysis. J. of Geotech. Eng., ASCE, 118(11), 1727-1747.
tailings.info, 2005. Mufulira Mine tailings breach, Zambia. On-line at
http://www.tailings.info/mufulira.htm.
Takahashi, H., Kitazume, M., Ishibasi, S., and Yamawaki, S. 2006. Evaluating the saturation
of model ground by P-wave velocity and modeling of models for a liquefaction
study. International Journal of Physical Modelling in Geotechnics 1, 13-15.
Taylor, D.W. 1948. Fundamentals of Soil Mechanics. John Wiley and Sons, Inc. New York.
Takahashi, H., Kitazume, M., Ishibasi, S., and Yamawaki, S. 2006. Evaluating the saturation
of model ground by P-wave velocity and modeling of models for a liquefaction
study. International Journal of Physical Modelling in Geotechnics 1, 13-15.
Tamura, S., Tokimatsu, K., Abe, A. and Sato, M. 2002. Effects of air bubbles on B-value
and P-wave velocity of a partly saturated sand. Soils and Foundations, 42(1), 121-
129.
Tsukamoto, Y., Ishihara, K., Nakazawa, H., Kamada, K. and Huang, H. 2002. Resistance of
partly saturated sand to liquefaction with reference to longitudinal and shear wave
velocities. Soils and Foundations, 42(6), 93-104.
Taiebat, M., 2008. Advanced eleastic-plastic constitutive and numerical modeling in
geomechanics. Ph.D. Thesis, University of California, Davis.
Taiebat, M., Jeremic, B., Dafalias, Y.F., Kaynia, A.M., and Cheng, Z., 2010. Propagation of
seismic waves through liquefied soils. Soil Dynamics and Earthquake Engineering,
30(4); pp. 236-257.
Taiebat, M., Jeremic, B., and Dafalias, Y.F., 2010. Prediction of seismically induced voids
and pore fluid volume/pressure redistribution in geotechnical earthquake
engineering. 63rd Canadian Geotechnical Conf., Canadian Geotechnical Society,
Edmonton, Alberta; pp. 233-237.
Terzaghi, K. And Peck R.B. 1967. Soil Mechanics in Engineering Practice. John Wiley and
Sons, Inc., New York.
Vaid, Y., Byrne, P.M., and Hughes, J.M.O. 1981. Dilation angle and liquefaction potential.
Journal of Geotechnical Engineering, ASCE, 103(7), 1003-1008.

166
References

Vaid, Y.P., and Chern, J.C. 1985. Cyclic and monotonic undrained response of saturated
sands. In Advances in the art of testing soils under cyclic conditions. Edited by V.
Khosla. ASCE Convention, Detroit, Mich. pp. 120147.
Vaid, Y.P., and Thomas, J. 1995. Liquefaction and post liquefaction behaviour of sand.
Journal of Geotechnical Engineering, ASCE, 121(2): 163173.
Vaid, Y.P., and Eliadorani, A., 1998. Instability and liquefaction of granular soils under
undrained and partially drained states. Canadian Geotech. J. Vol. 35, pp. 1053
1062.
Vasquez-Herrera, A. and Dobry, R., 1989. Re-evaluation of the Lower San Fernando Dam,
Rpt. 3. US Army Engineer Waterways Experiment Station, Contract Rpt. GL-89-2.
Whitman, R.V 1985. On liquefaction. In: Proceedings, 11th International Conference on
Soil Mechanics and Foundation Engineering. San Francisco, CA. A.A. Balkema, pp
1923-1926.
Wride, C. E., McRoberts, E. C., and Robertson, P. K. 1999. Reconsideration of case
histories for estimating undrained shear strength in sandy soils. Canadian
Geotechnical Journal, 36, 907-933.
Wu, J. 2002. Liquefaction triggering and post liquefaction deformations of monterey 0/30
sand under uni-directional cyclic simple shear loading. PhD Dissertation, University
of California, Berkeley, Calf. J. of Geotechnical Engineering, ASCE. Vol. 113(8):
861-878.
Yang, D., Naesgaard, E., Byrne, P.M. Adalier, K., and Abdoun, T. (2004). Numerical Model
Verification and Calibration of George Massey Tunnel Using Centrifuge Models.
Canadian Geot. J., 41(5), 921-942.
Yang. D., Naesgaard, E., Gohl, B., 2003. Geotechnical seismic retrofit design of immersed
George Massey Tunnel. 12th Panamerican Conf. on Soil Mechanics and
Geotechnical Engineering, June.
Yang, J., and Sato, T. 2000. Interpretation of seismic vertical amplification observed at an
array site. Bull. Seism. Soc. Am. 90, 275284.
Yang, J. 2002. Liquefaction resistance of sand in relation to P-wave velocity. Geotechnique,
52(4), 295-298.
Yang, J., Savidis, S. and Roemer, M. 2004. Evaluating liquefaction strength of partially
saturated sand. J. of Geotech. Geoenviron. Eng., ASCE, 130(9), 975-979.
Yang, J. 2005. Pore pressure coefficient for soil and rock and its relation to compressional
wave velocity. Geotechnique, 55(3), 251-256.
Yang, Z., and Elgamal, A., 2002. Influence of permeability on liquefaction-induced shear
deformation. Journal of Engineering Mechanics, ASCE, Vol. 128, No. 7, July, pp.
720-729.
Yoshida, N., and Finn, W.D.L., 2000. Simulation of liquefaction beneath an impermeable
surface layer. Soil Dyn. Earthquake Eng. 19, pp. 333-338.

167
References

Yoshimine, M., Robertson, P. K., and Wride, C. E. 1999. Undrained shear strength of clean
sands to trigger flow liquefaction. Canadian Geotechnical Journal, 36, 891-906.
Youd, T.L., Hansen, C.M., and Bartlett, S.F. 2002. Revised multilinear regression equations
for prediction of lateral spread displacement. J. Geotech. Geoenviron. Eng.,
128(12), 10071017.
Zienkiewicz, O.C. and Pastor, M. 1994. Computational mechanics and earthquake
engineering. Earthquake Engineering, Tenth World Conf., Balkema, Rotterdam,
ISBN 9054100605.

168
Appendix A

APPENDIX A

UBCSAND Flow Chart and Fish Code

UBCSAND1v02 FLAC Fish Code

NOTE:
Determining the appropriateness and accuracy of this routine for any purpose
is sole responsibility of end user. Routine is provided to specific
organizations by author and is not transferrable outside of this organization.
Please refer new users and potential bugs to primary author at
pmb@civil.ubc.ca.

******************************************************************************
* FISH version of UBCSAND MODEL from *
* Mohr-Coulomb model with *
* strain hardening/softening *
* Effective stress stress approach *
* primary and secondary plastic hardener *
* *
* Revisions *
* NOV 14 2001 pmb Change to post trigger plastic modulus and *
* crossover counter m_count4,m_ocr *
* DEC 27 2001 pmb m_triax = 1 to simulate comp ext tests *
* Feb 6 2002 pmb Modified plastic hardeners and basic relationship *
* between plastic and elastic moduli *
* Feb 13 2002 pmb Change to anisotropy (only for first time loading) *
* Sep 12 2002 mhb Change m_count4 to $gplim & $ratlim *
* modified $hard1 for m_n160 of 5 to 10 *
* modified m_dt at low $sig *
* reset 2ary yield surface if dilation *
* introduced zart for averaging stress components *
* limited maximum m_knew2 to m_knewp *
* March 30 2008 pmb Added one-sided loading and tension changes *
* Nov. 1 2008 PMB Changed start of Running section *
* Added m_mohr = 1.0 to simulate mohr model *
* Add m_hfac4 to allow reduction of dilation after triggering *
*
* January 8,2009 Modify one-sided loading *
* January 13,2009 Modify for drained condition or fmod = 0, use *
* m_hfac3 = 0 as a signal *
* May 12,2009 Modify for silt. Change m_hfac3 and m_hfac4 *
* July 24.2009 Modified $hard associated with m_hfac3 and $hard2 *

169
Appendix A

* associated with m_hfac4 *


* July26,2009 Changed default on $hard to 0.001. *
* Mar31-May12,2010 UBCSAND1v02 - General clean-up plus added: *
* -revised dilation cut-off option, *
* -revised pull-down of yield envelope on unloading, *
* -reduced default mean and vertical effective stress, *
* -revised tension yield envelope default, *
* -added zero dilation interval following passing *
* through zero stress origin, *
* -changed m_Pa to $Pa = atmospheric pressure, *
* -changed m_phicv to m_phipt *
* -changed m_ratcv to m_ratpt *
* (pt stands for phase transformation *
******************************************************************************
set echo off
def m_mss
constitutive_model 99
f_prop m_kge m_ne m_kb m_me m_ocr m_triax
f_prop m_kgp m_np m_phipt m_phif m_rf m_n160
f_prop m_g m_k m_coh m_ten m_ind
f_prop m_csnp m_nphi m_npsi m_e1 m_e2 m_x1 m_sh2
f_prop m_anisofac m_$fac m_css m_knew m_knew1 m_knew2
f_prop m_ratio m_ratpt m_ratf m_gpsum m_ratcrs m_knewp
f_prop m_dratmob m_ratmob m_dt m_flago m_ratmobold m_cross
f_prop m_hfac1 m_hfac2 m_hfac3 m_hfac4 m_hfac5 m_hfac6
f_prop m_epsum m_epsum1 m_rtymax m_ratmax m_ncyc m_ncyc1
f_prop m_epsav m_epsum4 m_epsum4old m_ratmax0 m_ratmax1 m_sxyold
f_prop m_ratioy m_rtmax m_gpstar m_mohr m_epsum5 m_epsum6
f_prop m_signal1 m_dilcut m_ratmax3 m_cnt m_epsum7 m_epsum8

float $sphi $spsi $s11i $s22i $s12i $s33i $sdif $s0 $rad
float $s1 $s2 $s3 $dc2 $dss
float $si $sii $psdif $fs $alams $ft $alamt $cs2 $si2
float $apex $epsav $tpsav $de1ps $de3ps $depm $eps $ept
float $bisc $pdiv $anphi $tco $sig $hard1 $area $hard2
float $sd $sxy $dumsig $dumsd $dumsxy $epn $epsum $cross
float $eps1 $epn1 $ratmax $hard $sy $dumsy $ratlim $gplim
float $epsum5 $Pa
;----------------------------------------------
Case_of mode
; ----------------------
; Initialisation section
; ----------------------
Case 1
; --- data check ---
$m_err = 0

170
Appendix A

if m_phif > 89.0 then


$m_err = 1
end_if
if m_coh < 0.0 then
$m_err = 3
end_if
if m_ten < 0.0 then
$m_err = 4
end_if
if $m_err # 0 then
nerr = 126
error = 1
end_if
; ----FLAG TO SET UP INITIAL CONDITIONS THE FIRST TIME IT GOES
THROUGH
;-----AND EACH RESTART
if m_flago < 5.0 then ;AVOIDS CHANGES ON RESTART
m_ratf = sin(m_phif * degrad)
m_ratpt = sin(m_phipt * degrad)
m_k = m_kb * $pa
m_g = m_kge * $pa
m_e1 = m_k + 4.0 * m_g / 3.0
m_e2 = m_k - 2.0 * m_g / 3.0
m_sh2 = 2.0 * m_g
; --- set tension to prism apex if larger than apex ---
$apex = m_ten
if m_phif # 0.0 then
$apex = m_coh / tan(m_phif * degrad)
end_if
m_ten = min($apex,m_ten)
end_if
if $ratlim = 0.0 then
$ratlim = 0.01 ;used for crossovers
end_if
if $gplim = 0.0 then
$gplim = 0.00005 ;used for crossovers
end_if
if m_n160 = 0.0 then
m_n160 = 5.0
end_if

Case 2
; ---------------
; Running section
; ---------------
m_flago = m_flago +1.0

171
Appendix A

if m_flago < 5.0 then ;FOR STARTUP


m_ratmob= m_ratf ;Treats as Mohr model
m_dt = m_ratpt - m_ratmob
$sphi = m_ratmob ; makes it elastic on startup
$spsi = -m_dt
m_npsi = (1.0 + $spsi) / (1.0 - $spsi)
m_nphi = (1.0 + $sphi) / (1.0 - $sphi)
m_x1 = m_e1 - m_e2*m_npsi + (m_e1*m_npsi - m_e2)*m_nphi
m_csnp = 2.0 * m_coh * sqrt(m_nphi)
if abs(m_x1) < 1e-6 * (abs(m_e1) + abs(m_e2)) then
$m_err = 5
nerr = 126
error = 1
end_if
end_if
zvisc = 1.0
if m_ind # 0.0 then
m_ind = 2.0
end_if
$anphi = m_nphi
; --- get new trial stresses from old, assuming elastic increments ---
$s11i = zs11 + (zde22 + zde33) * m_e2 + zde11 * m_e1
$s22i = zs22 + (zde11 + zde33) * m_e2 + zde22 * m_e1
$s12i = zs12 + zde12 * m_sh2
; $s33i = zs33 + (zde11 + zde22) * m_e2 + zde33 * m_e1
; $s33i = $s22i
$s33i = .5*($s11i+$s22i)
$sdif = $s11i - $s22i
$s0 = 0.5 * ($s11i + $s22i)
$rad = 0.5 * sqrt ($sdif*$sdif + 4.0 * $s12i*$s12i)
; ----principal stresses ---
$si = $s0 - $rad
$sii = $s0 + $rad
$psdif = $si - $sii
; --- determine case ---
; section
; if $s33i > $sii then
; --- s33 is major p.s. ---
; $icase = 3
; $s1 = $si
; $s2 = $sii
; $s3 = $s33i
; exit section
; end_if
; if $s33i < $si then
;; --- s33 is minor p.s. ---

172
Appendix A

; $icase = 2
; $s1 = $s33i
; $s2 = $si
; $s3 = $sii
; exit section
; end_if
; --- s33 is intermediate ---
$icase = 1
$s1 = $si
$s2 = $s33i
$s3 = $sii
; end_section
section
; --- shear yield criterion ---
$fs = $s1 - $s3 * $anphi + m_csnp
$alams = 0.0
; --- tensile yield criterion ---
$ft = m_ten - $s3
$alamt = 0.0
; --- tests for failure ---
if $ft < 0.0 then
$bisc = sqrt(1.0 + $anphi * $anphi) + $anphi
$pdiv = -$ft + ($s1 - $anphi * m_ten + m_csnp) * $bisc
if $pdiv < 0.0 then
; --- shear failure ---
$alams = $fs / m_x1
$s1 = $s1 - $alams * (m_e1 - m_e2 * m_npsi)
$s2 = $s2 - $alams * m_e2 * (1.0 - m_npsi)
$s3 = $s3 - $alams * (m_e2 - m_e1 * m_npsi)
m_ind = 1.0
else
; --- tension failure ---
$alamt = $ft / m_e1
$tco= $alamt * m_e2
$s1 = $s1 + $tco
$s2 = $s2 + $tco
$s3 = m_ten
m_ind = 3.0
; ------
end_if
else
if $fs < 0.0 then
; --- shear failure ---
$alams = $fs / m_x1
$s1 = $s1 - $alams * (m_e1 - m_e2 * m_npsi)
$s2 = $s2 - $alams * m_e2 * (1.0 - m_npsi)

173
Appendix A

$s3 = $s3 - $alams * (m_e2 - m_e1 * m_npsi)


m_ind = 1.0
else
; --- no failure ---
zs11 = $s11i
zs22 = $s22i
zs33 = $s33i
zs12 = $s12i
exit section
end_if
end_if
; --- direction cosines ---
if $psdif = 0.0 then
$cs2 = 1.0
$si2 = 0.0
else
$cs2 = $sdif / $psdif
$si2 = 2.0 * $s12i / $psdif
end_if
; --- resolve back to global axes ---
case_of $icase
case 1
$dc2 = ($s1 - $s3) * $cs2
$dss = $s1 + $s3
zs11 = 0.5 * ($dss + $dc2)
zs22 = 0.5 * ($dss - $dc2)
zs12 = 0.5 * ($s1 - $s3) * $si2
zs33 = $s2
case 2
$dc2 = ($s2 - $s3) * $cs2
$dss = $s2 + $s3
zs11 = 0.5 * ($dss + $dc2)
zs22 = 0.5 * ($dss - $dc2)
zs12 = 0.5 * ($s2 - $s3) * $si2
zs33 = $s1
case 3
$dc2 = ($s1 - $s2) *$cs2
$dss = $s1 + $s2
zs11 = 0.5 * ($dss + $dc2)
zs22 = 0.5 * ($dss - $dc2)
zs12 = 0.5 * ($s1 - $s2) * $si2
zs33 = $s3
end_case
; zvisc = 0.0
end_section

174
Appendix A

; -----------------------------------------------------------------------
; -----UBC add on to account for change of elastic and plastic parameters
; -----------------------------------------------------------------------
; -------- PLASTIC STRAINS ---
if m_ind = 1.0 then
$de1ps = $alams
$de3ps = -$alams * m_npsi
$eps1 = abs($de1ps-$de3ps)
$epn1 = -($de1ps+$de3ps)
$eps = $eps + $eps1*zart ; always positive
$epn = $epn + $epn1*zart
end_if
;-------------- STRESSES
$sig = -0.5*(zs11+zs22)
$sd = -(zs11-zs22) / 2.0
$sxy = zs12
$sy = -zs22
$dumsig = $dumsig + $sig*zart
$dumsd = $dumsd + $sd *zart
$dumsxy = $dumsxy + $sxy*zart
$dumsy = $dumsy + $sy *zart
$area = $area + zart
SECTION
; ---GET AVERAGE VALUES OF STRESSES AND STRAINS
if zsub > 0.0 then ;zsub loop
;-----STRAINS
$epsav = 0.0
$epsum = 0.0
$epsav = $eps / $area ;PLASTIC SHEAR STRAIN INCREMENT
$epsum = $epn / $area ;PLASTIC VOLUMETRIC STRAIN INCREMENT
$eps = 0.0
$epn = 0.0
;---- STRESSES
$sig = $dumsig/$area
$sig = max($sig,0.0005*$pa) ;0.005 APR 5 EN DECREASE DEFAULT BY 10
$sd = $dumsd/$area/$sig
$sxy = $dumsxy/$area/$sig
$sy = $dumsy/$area
$sy = max($sy,0.0005*$pa) ;0.005 APR 5 EN DECREASE DEFAULT BY 10
m_ratioy = $sxy/$sy * $sig
if m_triax = 1.0 then ;To simulate triaxial or plane strain tests having sxy
= 0 only
m_ratioy = (1.0 - $sy/$sig) ;When sxy = 0.0, control by sxx-syy
end_if
m_ratio = sqrt($sd*$sd+$sxy*$sxy)
m_ratio = min(m_ratio,m_ratf)

175
Appendix A

if abs(m_ratioy)> m_ratf then


m_ratioy = m_ratioy*m_ratf/abs(m_ratioy) ;EN Apr 06 2010
end_if
$dumsd = 0.0
$dumsxy = 0.0
$dumsig = 0.0
$dumsy = 0.0
$area = 0.0
;************************************
; Resets yield loci and other factors
if m_ratmax = 0. then
m_ratmax3 = 0.0 ; maximum mobillized stress ratio when
; cycles do not have cross-over. Reset to
; zero each time there is a cross-over.
m_ratmax1 = 0.0
m_ratmax0 = 0.0
m_ratmax = 1.0
m_ratmob = m_ratio
m_epsum = 0.0
m_epsum1 = 0.0
m_epsum4 = 0.0
m_epsum5 = 0.0
m_epsum6 = 0.0
m_ncyc = 0.0
m_ncyc1 = 0.0
m_ratcrs = m_ratio
m_cross = 10.0
m_signal1 = 0.0
end_if
;**************************************

m_dratmob = 0.0
if $epsav > 0.0 then ; PLASTIC LOOP
m_epsum = m_epsum + $epsum
m_epsav = m_epsav + $epsav
m_epsum7 = m_epsum7 + $epsav ; EN Apr 8 2010 cumulated plastic shear
strain following cross-over

if $epsum < 0.0 then ; DILATION


m_epsum1 = m_epsum1 + $epsav ; Nov1 2008 accumulated Plastic shear
; strain associated with expansion.
; Set to ;zero at end of each half
; cycle
m_epsum5 = m_epsum5 + $epsav ; May 11, 2009 accumulated Plastic
; shear strain associated with
; expansion.

176
Appendix A

end_if
m_ratmax0 = max(m_ratmax0,m_ratioy)
m_ratmax1 = min(m_ratmax1,m_ratioy)

;----Evaluate anisotropy factor


m_css=$cs2
if m_css>=0.0 then
m_$fac = m_anisofac
else
m_$fac = m_anisofac + (m_anisofac - 1.0) * m_css
end_if

;----PLASTIC SHEAR MODULUS


m_knew = m_kgp/$sig * $pa*($sig/$pa)^m_np*m_hfac1

;-------secondary yield:
$hard1 = max(0.5, 0.1*m_n160) ;correction at low N160
$hard1 = min(1.0, $hard1)
m_knew1 = m_knew*( 4. + m_ncyc1) *$hard1 * m_hfac2

;-------primary yield:
if m_ocr <= 2.0 then
if m_ratioy > 0.0 then
$ratmax = m_ratmax0
else
$ratmax = abs(m_ratmax1)
end_if
if abs(m_ratioy) > 0.99*$ratmax then
m_knew1 = 0.5*(m_knew+m_knew1)*m_$fac
end_if
end_if
;-------modify for stress ratio:
m_gpstar = m_knew1 *(1.0-(m_ratio*m_rf/m_ratf))^2
; Modify for dilation
;-------dilation "softener" to control post-liq:
;-------m_epsum4 is the accumulated shear strain associated with dilation during
; the previous stress pulse.
$hard = 1.0
if m_hfac3 > 0.0
if m_epsum4 > 1e-6 then ;
if m_signal1 = 0.0 ; added July 24,2009
$epsum5 = m_epsum5
else
$epsum5 = m_epsum6
end_if

177
Appendix A

$hard = max(0.001,exp(-$epsum5*110.0*m_hfac3)) ; $hard range


; 1 to 0.001,
; July 26,09.

;;; m_knew2 = m_knew * $hard ; add this line if want to plot m_knew2
m_gpstar = m_knew * $hard

end_if
end_if
;--- Raise Yield locus, m_ratmob
m_dratmob = m_gpstar*$epsav
m_dratmob = max(m_dratmob,0.0)
m_ratmob = m_ratmob + m_dratmob
m_ratmob = min(m_ratmob,m_ratf) ;current yield locus
end_if
;------------------------------------ ;END OF PLASTIC LOOP

;---- Lower m_ratmob upon unloading*************************************


if $epsav = 0.0 ; All sub zones elastic

If m_ratio < 0.90 * m_ratmob


m_ratmax3 = max(m_ratmax3,m_ratmob) ; EN Apr 4 2010
if m_ratmax3 < m_ratpt then ; EN Apr 4 2010
m_ratmob = 0.95*m_ratmob
else
if m_cnt < 1.0 then
m_ratmob = max(0.95*m_ratmob,(m_ratpt*m_hfac6)) ; EN April 3 2010
end_if
end_if
end_if

end_if

;---CROSSING AXIS RESETS PLASTIC PARAMETERS


if m_ratioy*m_sxyold < 0.0 then ;crossover check

;------Crossover has occurred


$cross = max(m_rtymax,m_rtmax-m_ratcrs) / $ratlim
$cross = max($cross, m_gpsum/$gplim)
m_ratcrs = m_ratio
m_rtmax = m_ratio
m_gpsum = 0.0
m_rtymax = 0.0

;-------Previous half cycle is "large"

178
Appendix A

if max(m_cross, $cross) > 1.0 then


if m_cross # 99.0 then
m_ncyc = m_ncyc + 0.5
m_ncyc1 = m_ncyc1 + 0.5
m_ratmobold = m_ratmob
else
m_ratmobold = max(m_ratmob,m_ratmobold)
endif
m_ratmob = m_ratio
m_ratmax3 = 0.0 ; EN Apr 4 2010
m_epsum4old = m_epsum4
m_epsum4 = m_epsum1 ; preserves the prior dilation

if m_epsum4 > 1e-6 then ; reset 2ary yield surface if dilation


m_ratmax0 = 0.0
m_ratmax1 = 0.0
m_ncyc1 = 0.0
endif

m_epsum1 = 0.0
if m_cnt < 1.0 then
m_epsum8 = m_epsum7 ; comment 1000
m_epsum7 = 0.0 ;EN Apr 8 2010
end_if

m_cross = 0.0

;-------Previous half cycle is "small"


else
m_ratmob = m_ratio + 0.75*(max(m_ratmobold,m_ratio) - m_ratio)
m_epsum1 = m_epsum4
m_epsum4 = m_epsum4old
m_cross = 99. ;remember small half cycle
endif
else

;-------No crossover
m_gpsum = m_gpsum + $epsav ;ignore initial crossover step (uses
old parameters)
m_rtymax = max(m_rtymax, abs(m_ratioy))
m_rtmax = max(m_rtmax, m_ratio)
end_if ; End Crossing loop

m_sxyold = m_ratioy

179
Appendix A

;---COMPUTE NEW PARAMETERS ACCORDING TO THE CURRENT MOBILIZED


FRIC ANGLE
;---------- SET PLASTIC VALUES
if m_mohr = 1.0
m_ratmob = m_ratf
end_if

m_dt = m_ratpt-m_ratmob
m_dt = min(m_dt,0.5*m_ratpt) ; EN April 2010
if m_epsum4 > 1e-6 ;Dilation occurring
if m_dt > 0.0 then
m_dt=(m_ratpt-m_ratmob)
else
; Reduce Dilation with plastic shear strain
if m_signal1 = 1.0
$epsum5 = m_epsum5- m_epsum6
$hard2 = max(0.01, exp(-($epsum5)^.25*m_hfac4)) ; 0.01 to 0.1
m_dt=(m_ratpt-m_ratmob)*($hard2)
end_if
end_if
end_if

if $sig < 0.01*$pa then


m_cnt = 99.0
if m_epsum7 < m_epsum8*m_hfac5 then ; if plastic strain since crossover <
function cummulated plastic dilation strain then
m_ratmob = m_ratpt
else
m_ratmob = m_ratf ; added March 23 2010
end_if

m_dt=(m_ratpt-m_ratmob)

if m_signal1 = 0.0 ; Added July 24,2009


m_epsum6 = m_epsum5
m_signal1 = 1.0
end_if
else
m_cnt = 0.0
end_if
m_ratmob = max(m_ratmob,0.01)

; Dilation cut-off (m_dilcut controlled outside of constitive model)


if m_dilcut = 1.0 then
m_dt = max(m_dt,0.0)
end_if

180
Appendix A

;---------- PLASTIC PARAMETERS


$sphi = m_ratmob
$spsi = -m_dt
m_npsi = (1.0 + $spsi) / (1.0 - $spsi)
m_nphi = (1.0 + $sphi) / (1.0 - $sphi)

; ---STRESS DEPENDENT ELASTIC MODULI

m_k = m_kb * $pa * ($sig/$pa)^m_me


m_g = m_kge * $pa * ($sig/$pa)^m_ne
m_e1 = m_k + 4.0 * m_g / 3.0
m_e2 = m_k - 2.0 * m_g / 3.0
m_sh2 = 2.0 * m_g
m_x1 = m_e1 - m_e2*m_npsi + (m_e1*m_npsi - m_e2)*m_nphi
end_if ;--------------------------------------END OF ZSUB > 0
END_SECTION

Case 3
; ----------------------
; Return maximum modulus
; ----------------------
if m_g = 0.0 then
m_k = m_kb * $pa
m_g = m_kge * $pa
end_if
cm_max = (m_k + 4.0 * m_g / 3.0)
sm_max = m_g

Case 4
; ---------------------
; Add thermal stresses
; ---------------------
ztsa = ztea * m_k
ztsb = zteb * m_k
ztsc = ztec * m_k
ztsd = zted * m_k
End_case
end
opt m_mss
set echo=on

181
Appendix A

Typical FLAC fish properties code for UBCSAND1v02

def properties
loop i (1,izones)
loop j (1,jzones)
;ELASTIC
m_n160(i,j) = max(m_n160(i,j),1.0)
m_kge(i,j) = 21.7*25.*m_n160(i,j)^.333 ; Max. Normalized shear Shear Mod
m_kb(i,j) = m_kge(i,j)*0.916 ; Bulk mod = m_kge* (2(1+mu))/(3*(1-2*mu))
m_me(i,j) = 0.5 ;mu = 0.1 , factor = 0.916
m_ne(i,j) = 0.5
;PLASTIC PROPERTIES
m_kgp(i,j) = m_kge(i,j)* m_n160^2*.003 +100.0 ;shear Mod
m_np(i,j) = 0.4

m_phipt(i,j)= min(33.0,40.0-m_n160(1,1)*0.75)
m_phif(i,j) = 33.0 + (m_n160(1,1)/10.)^1.65

;plastic modification factors

m_hfac1(i,j) = 0.65 ;primary hardener, Controls no. of cycles to trigger liquefaction


m_hfac2(i,j) = 0.85 ;Secondary hardener, Refines shape of pore pressure rise with
;cycles
m_hfac3(i,j) = 1.0 ;dilation "hardener", Controls post-trigger response
m_hfac4(i,j) = 0.6 ;reduces dilation after triggering, 0 = no reduction, 1 is large
;reduction
m_hfac5(i,j) = 1.0 ;set length of zero dilation zone following going to origin
m_hfac6(i,j) = 0.95 ;set amount of failure envelope pulldown below phipt on
;unloading

;failure ratio --same as in Hyperbolic model


m_rf(i,j) = 1.0 - m_n160(i,j)/100.
m_rf(i,j) = max(m_rf(i,j),.5)
m_rf(i,j) = min(m_rf(i,j),.99)
;plastic anisotrophy
;m_anisofac(i,j) = .0166*m_n160(i,j)
;m_anisofac(i,j) = min(m_anisofac(i,j),1.0)
;m_anisofac(i,j) = max(m_anisofac(i,j),0.333)
m_anisofac(i,j) = 1.0
;m_anisofac ;Anisotrophy factor; 1 for isotropic, .333 for loose pluviated
end_loop
end_loop
end

182
Appendix A

Figure A-1 Flow diagram for UBCSAND1v02 illustrating how it fits within the FLAC
(ITASCA 2008) Mohr Coulomb framework.

183
Appendix B

APPENDIX B

ANALYSIS OF A CONCRETE GRAVITY DAM OVER POTENTIAL


LIQUIFIABLE SOIL ILLUSTRATING PROPOSED HYBRID
PROCEDURE METHODOLOGY

B.1 Introduction

This example project is given to illustrate the design procedure that has been developed using
UBCSAND and UBCHYST with a post-shaking total stress check for stability with
empirical post-liquefaction residual strengths. The analyses were carried out for a concrete
gravity dam over potentially liquefiable foundation soils. This is work from an actual project
but is given here as an example analysis using the proposed hybrid effective stress total
stress procedure.

The structure analyzed consisted of a concrete intake structure with downstream


penstocks founded on fluvial, deltaic and glacial sediments. The analyses were carried out to
assess the seismic performance of the concrete intake structure on a section normal to the
axis of the dam.

The intake structure is a concrete gravity dam approximately 13m high, 43m wide
by 24m long, and has a sheet-pile and jet-grout installed bentonite-cement cut-off wall
located under and adjacent to the structure.

The dam is located in an area of high seismicity. Preliminary analyses were carried
out using seven natural earthquake records that were linearly scaled so that the spectral
values of the average response spectrum are similar to those of the uniform hazard response
spectrum within the natural period of interest between 0.75s and 1.5s. Peak ground
acceleration of the records varied from 0.52 to 0.74 g.

B.2 Dynamic Numerical Analyses

Two dimensional (2-D) non-linear dynamic numerical analyses were carried out using the
finite difference program FLAC version 6.0 (ITASCA 2008). The analyses were carried out
in ground water mode and flow and pore pressure redistribution were allowed.

184
Appendix B

Cohesionless (sandy) soils were modeled using the effective-stress constitutive model
UBCSAND, while non-liquefiable cohesive (clay/silt) soils were modeled using the total
stress constitutive model UBCHYST. In this context, effective stress refers to constitutive
models where shear strain, skeleton volume change, and pore pressure are coupled and
directly included in the model. In the total stress model, shear strain does not induce
volume or related pore pressure change; however, pore pressure changes are indirectly
accounted for by reducing stiffness and strength when pore pressure build-up or liquefaction
is predicted. The concrete intake structure and reservoir water were included in the model.

Figure B-1 shows a typical numerical model profile and material zoning.
Descriptions of material properties are summarized in Table B-1. The model grid is
illustrated in Figure B-2.

In FLAC, the dynamic analyses were carried out in groundwater mode in a


chronological manner simulating in-situ conditions as described below.

The general procedure used for analyses included the following steps:

Set up model grid, material properties and pore water regime and bring to
static equilibrium using Elastic and then Mohr Coulomb constitutive models.

Select representative elements for calibration of UBCSAND and UBCHYST.


For each selected element note the representative material properties (shear
wave velocity, vertical and lateral effective stress, and (N 1 ) 60 values) for use
in the calibration.

Calibrate UBCSAND constitutive model by exercising a single element


model. Calibration factors are adjusted so the response is in agreement with
the Idriss and Boulanger (2008) liquefaction triggering chart and to match
typical post-liquefaction response from laboratory tests.

Calibrate UBCHYST to give similar cyclic softening to that inferred from the
project cyclic triaxial laboratory tests.

Enter calibration factors in the large model and bring to static equilibrium.

Turn on flow and change to dynamic configuration with large strain, multi-
stepping, and nominal 1% Rayleigh damping and bring to equilibrium by
185
Appendix B

running with input motion of zero. (When damping, model parameters, or


FLAC configuration (i.e., static to dynamic mode) are changed in FLAC it
often takes time for the model to return to equilibrium. Because of this, it has
been found to be good practice to let the model come to equilibrium in
dynamic mode (by running with zero input motion) prior to carrying out the
dynamic analysis).

Firm ground (Pleistocene soil) outcropping time histories were converted to a


format that could be used for input to the base of the FLAC numerical model.
Two alternative procedures were used:

1. The half-space below the model was assumed to be flexible and input
motions were given as stress time histories; or,
2. The half-space was assumed rigid and input motions were given as
velocity time histories.
Set displacements to zero, apply earthquake time history at base, and solve
past end of earthquake.
Check model resistance against static flow slide using empirically developed
Idriss and Boulanger (2008) residual strengths (Figure 4-1(d)).
Calculate additional post-liquefaction consolidation settlement using
equations by Ishihara and Yoshimine (as given in Idriss and Boulanger 2008).
Compile and summarize results of analyses.

B.3 Results

Typical end-of-earthquake shaking displacements are shown in Figures B-3 and B-4 for the
horizontal and vertical directions respectively. Typical pore pressure ratio (R u ) at the end of
earthquake shaking is shown in Figure B-5 and displacement time histories are shown on
Figure B-6. As shown in Figure B-6, changing to the empirical residual soil strengths (based
on back-analyses of case histories) in all zones that liquefied (R u > 0.7) did not cause any
significant additional movements.

186
Appendix B

B.4 Application Example Discussion and Conclusions

Over 50 dynamic coupled effective stress analyses were carried out with various soil
parameters and model profiles. Select analyses have been checked for post-liquefaction
stability using empirically derived residual strengths within liquefied zones. Select analyses
to estimate post-shaking consolidation settlement were also carried out. Due to the
complexity of the problem and approximations in the methodology there will be significant
uncertainty in the calculated displacements. However, the analyses give an approximate
range of displacements and give considerable insight into potential behaviour and failure
modes.

Key findings of the analyses are as follows:

(1) the intake structure only moves during earthquake shaking and stops moving at end
of shaking (excepting post-shaking consolidation settlement);
(2) the main movements are of a downward bearing failure type and a lateral shifting
that is resulting from shear in relatively deep liquefied zones;
(3) there is some tilting of the intake structure (upstream end settles more than
downstream end);
(4) there will be a relatively sharp differential vertical displacement (approximately.
0.1m to 0.7m) between the downstream toe of the intake structure and the ground
surface a few metres further downstream; and,
(5) depending on the earthquake record and model profile, the intake structure may
move either upstream or downstream. In some analyses, reversing the direction of
earthquake motion resulted in a change in the direction of net lateral movement.

A post-liquefaction stability check in which the shear strength of liquefied zones


were set to an empirically derived residual strength by Idriss and Boulanger (2008) did not
result in any significant movements of the dam. Reducing the residual strength by a factor of
1.2 did not significantly change the result.

In all the analyses, lateral movement of the dam centre was less than 0.6m
(upstream or downstream) and shear strain settlement was less than 0.9m. Post-shaking
consolidation settlements were all less than 0.2m. In analyses where the upper silt (Unit 3a
and 3b in Figure B-1) was assumed to behave as loose sand, shear strain settlements were

187
Appendix B

more than double and post-shaking consolidation settlements were more than fifty percent
higher than when the silt was assumed to cyclically strain soften but not liquefy in a loose
sand-like manner.

Of the seven final earthquakes run, the Chi Chi TCU071-W earthquake record gives
the highest settlements. For identical model and assumptions, the TCU071-W record gives
settlements that are 50% higher than the average settlements (TCU071-W gives -0.75m,
whereas, the average is -0.5m).

Including vertical motion in addition to the horizontal base, excitation did not
significantly change the resulting displacements. Applying the input motions as a velocity in
lieu of as a stress history appeared to give slightly higher vertical settlements (approximately
10%).

Calculated displacements were similar whether the water in the reservoir was
modeled as an extremely weak soil or as applied pressures to the mud-line and dam.

Extending the downstream length of the mesh by 100m resulted in a small (10%)
decrease in settlement.

In all analyses, lateral movements at the center of the intake structure were less than
about 0.5m, vertical shear induced settlement less than 0.9m and post-shaking consolidation
settlements of the intake structure less than 0.2m. Some analyses were carried out assuming
that the upper silt soil units 3a and 3b would behave as loose sand and are deemed to give
conservative upper-bound displacements. Considering this, best estimate centre of dam
displacements, including post-shaking consolidation settlement, are deemed to be about 0.5m
lateral movement (could be either upstream or downstream) and 0.5m settlement. On
average, the upstream toe of the intake structure settled approximately 25% more and the
downstream toe 33% less than the centre of the dam. Considering the various uncertainties,
values in the range one-half to double the best estimate lateral displacement values and one-
half to one and one half (1.5x) the vertical settlements are suggested for the assessment of the
intake structure.

188
Appendix B

Table B-1 Summary of Soil Properties used in Example Analysis of Appendix B

SOIL ELEVATION(1) CTR(2) (N1)60-CS POROSITY DRY FRICTION PERMEABILITY


(3) (4)
LAYER TYPE TOP BOT THICKNESS DEPTH DYN RESID (n) DENSITY ANGLE(5) COHESION Vs k k(flac) MODEL
(m) (m) (m) (m) (blows/ft) (blows/ft) (kg/m^3) (degrees) (kPa) (m/s) (cm/s) (m s/kg)
1b SAND 133 127 6 3 25 25 0.39 1632 38 0 300 0.00042 4.2E-10 UBCSAND
2a SAND/SILT 127 122 5 8.5 23 21 0.44 1576 0 0 300 1.6E-06 1.6E-12 UBCSAND
3a(M) SILT 122 119 3 12.5 14 N/A 0.39 1545 34.1 145 300 0.000012 1.2E-11 UBCHYST
3a(S) SILT 122 119 3 12.5 14 14 0.39 1545 34.1 0 300 0.000012 1.2E-11 UBCSAND
2b SAND 119 117 2 15 20 20 0.44 1633 0 0 300 1.6E-06 1.6E-12 UBCSAND
3b(M) SILT 117 112 5 18.5 14 N/A 0.41 1550 0 145 300 1.6E-06 1.6E-12 UBCHYST
3b(S) SILT 117 112 5 18.5 14 14 0.41 1550 34.1 0 300 1.6E-06 1.6E-12 UBCSAND
2c SAND/SILT 112 106 6 24 22 20 0.41 1574 35.2 0 300 1.6E-06 1.6E-12 UBCSAND
4 SAND 106 99.5 6.5 30.25 34 N/A 0.39 1639 36.4 0 330 0.000012 1.2E-11 UBCSAND
5 SILT 99.5 79 20.5 43.75 10 N/A 0.41 1585 0 145 310 1.6E-06 1.6E-12 UBCHYST
TILL SAND/SILT 79 77 2 55 N/A 0.41 1650 N/A N/A 760 1.6E-06 1.6E-12 ELASTIC
J JET GROUT N/A 0.51 1300 0 100 275 1E-07 1E-13 UBCHYST
m BOTTOM MUD 0.55 1243 0 0.5 40 1.6E-06 1.6E-12 UBCHYST
W WATER 1000 0.1 to 1.0 0.1 to 1.0 N/A MOHR
Notes:
(1) ELEVATION TAKEN IN SECTION IMMEDIATELY DOWNSTREAM OF INTAKE STRUCTURE (at i =185 (x=20 m))
(2) CTR DEPTH = Depth of center of layer in section by downstrean end of intake structure
(3) DYN = equivalent clean sand (N1)60-CS for use with UBCSAND in dynamic analyses
(4) RESID = equivalent clean sand (N1)60-CS for use in post-shaking residual strength calculation
(5) FRICTION ANGLE DURING DYNAMIC ANALYSIS = 33 + (N1)60/10

Figure B-1 Example Dam analysis model profile and soils types. Soil properties are summarized
in Table B-1.

189
Appendix B

Figure B-2 Example Dam analyses model grid details.

Figure B-3 Example Dam analysis typical post-earthquake horizontal displacement contours.

190
Appendix B

Figure B-4 Example Dam analyses showing typical post-earthquake vertical displacement
contours.

Figure B-5 Example Dam analysis showing typical post-earthquake pore pressure ratio contours.

191
Appendix B

Figure B-6 Time histories showing displacement of center of intake structure and model base (top
of till) for typical analysis (positive is downstream or upward).

192

You might also like