Vous êtes sur la page 1sur 297

State of the Art and Practice in the

Assessment of Earthquake-Induced Soil


Liquefaction and Its Consequences

Committee on State of the Art and Practice in Earthquake Induced Soil


Liquefaction Assessment

Board on Earth Sciences and Resources

Division on Earth and Life Studies


THE NATIONAL ACADEMIES PRESS 500 Fifth Street, NW Washington, DC 20001

This activity was supported by a grant from the American Society of Civil Engineers, an award
from the Los Angeles Department of Water and Power, an award from the Port of Long Beach,
an award from the Port of Los Angeles, Grant No. R11AP81544 from U.S. Bureau of
Reclamation, Award No. DTFH6114P00075 from the U.S. Department of Transportation’s
Federal Highway Administration, and Award No. NRC-HQ-12-G-04-0061 with the U.S. Nuclear
Regulatory Commission. Any opinions, findings, conclusions, or recommendations expressed in
this publication do not necessarily reflect the views of any organization or agency that provided
support for the project.
ISBN-10: 0-309-44027-0

ISBN-13: 978-0-309-44027-1

Digital Object Identifier: 10.177226/23474


Additional copies of this publication are available for sale from the National Academies
Press, 500 Fifth Street, NW, Keck 360, Washington, DC 20001; (800) 624-6242 or (202)
334-3313; http://www.nap.edu.

Copyright 2016 by the National Academy of Sciences. All rights reserved.

Printed in the United States of America

Suggested citation: National Academies of Sciences, Engineering, and Medicine. 2016. State
of the Art and Practice in the Assessment of Earthquake-Induced Soil Liquefaction and
Its Consequences. Washington, DC: The National Academies Press. doi: 1017226/23474.

Спизжено у ExLib: avxhome.in/blogs/exLib


Stole src from http://avxhome.in/blogs/exLib:
tanas.olesya (avax); Snorgared, D3pZ4i & bhgvld, Denixxx (for softarchive)
My gift to leosan (==leonadin GasGeo&BioMedLover from ru-board :-) - Lover to steal and edit
someone else's
Любителю пиздить и редактировать чужое
The National Academy of Sciences was established in 1863 by an Act of Congress, signed by
President Lincoln, as a private, nongovernmental institution to advise the nation on issues
related to science and technology. Members are elected by their peers for outstanding
contributions to research. Dr. Marcia McNutt is president.

The National Academy of Engineering was established in 1964 under the charter of the
National Academy of Sciences to bring the practices of engineering to advising the nation.
Members are elected by their peers for extraordinary contributions to engineering. Dr. C. D.
Mote, Jr., is president.

The National Academy of Medicine (formerly the Institute of Medicine) was established in
1970 under the charter of the National Academy of Sciences to advise the nation on medical
and health issues. Members are elected by their peers for distinguished contributions to
medicine and health. Dr. Victor J. Dzau is president.

The three Academies work together as the National Academies of Sciences, Engineering,
and Medicine to provide independent, objective analysis and advice to the nation and
conduct other activities to solve complex problems and inform public policy decisions. The
National Academies also encourage education and research, recognize outstanding
contributions to knowledge, and increase public understanding in matters of science,
engineering, and medicine.

Learn more about the National Academies of Sciences, Engineering, and Medicine at
www.national-academies.org.
Reports document the evidence-based consensus of an authoring committee of experts.
Reports typically include findings, conclusions, and recommendations based on information
gathered by the committee and committee deliberations. Reports are peer reviewed and are
approved by the National Academies of Sciences, Engineering, and Medicine.

Proceedings chronicle the presentations and discussions at a workshop, symposium, or other


convening event. The statements and opinions contained in proceedings are those of the
participants and have not been endorsed by other participants, the planning committee, or
the National Academies of Sciences, Engineering, and Medicine.

For information about other products and activities of the National Academies, please visit
nationalacademies.org/whatwedo.
COMMITTEE ON STATE OF THE ART AND PRACTICE IN EARTHQUAKE INDUCED
SOIL LIQUEFACTION ASSESSMENT

EDWARD KAVAZANJIAN, JR., Chair, Arizona State University, Tempe


JOSÉ E. ANDRADE, California Institute of Technology, Pasadena
KANDIAH “ARUL” ARULMOLI, Earth Mechanics, Inc., Fountain Valley, California
BRIAN F. ATWATER, U.S. Geological Survey and University of Washington, Seattle
JOHN T. CHRISTIAN, Independent Consultant, Burlington, Massachusetts
RUSSELL GREEN, Virginia Polytechnic Institute and State University, Blacksburg
STEVEN L. KRAMER, University of Washington, Seattle
LELIO MEJIA, AECOM, Oakland, California
JAMES K. MITCHELL, Virginia Polytechnic Institute and State University (Retired), Blacksburg
ELLEN RATHJE, The University of Texas at Austin
JAMES R. RICE, Harvard University, Cambridge, Massachusetts
YUMEI WANG, Oregon Department of Geology and Mineral Industries, Portland

THE NATIONAL ACADEMIES OF SCIENCES, ENGINEERING, AND MEDICINE STAFF

SAMMANTHA MAGSINO, Study Director


COURTNEY R. GIBBS, Program Associate

v
BOARD ON EARTH SCIENCES AND RESOURCES

GENE WHITNEY, Chair, Congressional Research Service (Retired), Washington, DC


R. LYNDON (LYN) ARSCOTT, International Association of Oil & Gas Producers (Retired), Danville,
California
CHRISTOPHER (SCOTT) CAMERON, GeoLogical Consulting, LLC, Houston, Texas
CAROL P. HARDEN, The University of Tennessee, Knoxville
T. MARK HARRISON, University of California, Los Angeles
ANN S. MAEST, Buka Environmental, Boulder, Colorado
DAVID R. MAIDMENT, The University of Texas at Austin
M. MEGHAN MILLER, UNAVCO, Inc., Boulder, Colorado
ISABEL P. MONTAÑEZ, University of California, Davis
HENRY N. POLLACK, University of Michigan, Ann Arbor
MARY M. POULTON, University of Arizona, Tucson
JAMES M. ROBERTSON, Wisconsin Geological and Natural History Survey, Madison
SHAOWEN WANG, University of Illinois at Urbana-Champaign

THE NATIONAL ACADEMIES OF SCIENCES, ENGINEERING, AND MEDICINE STAFF

ELIZABETH A. EIDE, Director


ANNE M. LINN, Scholar
DEBORAH GLICKSON, Senior Program Officer
SAMMANTHA L. MAGSINO, Senior Program Officer
NICHOLAS D. ROGERS, Financial and Research Associate
COURTNEY R. GIBBS, Program Associate
ERIC J. EDKIN, Senior Program Assistant
RAYMOND M. CHAPPETTA, Program Assistant

vi
Acknowledgments

The committee relied on input from the geotechnical engineering community in addition to
the expertise contained within the itself. Thoughtful input from many technical experts informed
committee deliberations. We gratefully acknowledge the following individuals for giving
presentations or serving on panels during the committee’s two-day community workshop or
during information-gathering sessions at committee meetings: Donald Anderson, CH2M Hill;
Ronald Andrus, Clemson University; Pedro Arduino, University of Washington; Gregory
Baecher, University of Maryland; Laurie Baise, Tufts University; Steven Bartlett, University of
Utah; Michael Beaty, Beaty Engineering, LLC; Ronaldo Borja, Stanford University; Scott
Brandenberg, University of California, Los Angeles; Giuseppe Buscarnera, Northwestern
University; Misko Cubrinovski, University of Canterbury, New Zealand; Craig Davis, Los
Angeles Department of Water and Power; Ricardo Dobry, Rensselaer Polytechnic Institute;
Leslie Harder, HDR, Inc.; David Gillette, U.S. Bureau of Reclamation; Antonio Gioiello, Port of
Los Angeles; Thomas Holzer, U.S. Geological Survey; Kenji Ishihara, Chuo University, Tokyo;
Richard Iverson, U.S. Geological Survey; Robert Kayen, U.S. Geological Survey; Geoffrey
Martin, University of Southern California; Milan Pavich, U.S. Geological Survey; Peter
Robertson, Gregg Drilling and Testing, Inc.; Robert Schweinfurth, Geo-Institute/American
Society of Civil Engineers; Jonathan Stewart, University of California, Los Angeles; Thomas
Weaver, U.S. Nuclear Regulatory Commission; and T. Les Youd, Brigham Young University
(Emeritus).
Ross Boulanger, University of California, Davis; K. Önder Çetin; Middle East Technical
University, Turkey; Izzat M. Idriss, University of California, Davis; and Raymond Seed,
University of California, Berkeley, provided helpful written responses to questions from the
committee.
The committee also thanks Tarek Abdoun, Rensselaer Polytechnic Institute; Scott Anderson,
U.S. Federal Highway Administration; Christopher Baxter, University of Rhode Island; Brady
Cox, The University of Texas at Austin; Yannis Dafalias, University of California, Davis; Shideh
Dashti, University of Colorado Boulder; Roupen Donikian, Parsons Brinckerhoff, Inc.; Kevin
Franke, Brigham Young University; Ian Friedland, U.S. Federal Highway Administration;
Daniel Gillins, Oregon State University; Youssef Hashash, University of Illinois; Jianping Hu,
Los Angeles Department of Water and Power; Harold Magistrale, FM Global; Majid Manzari,
George Washington University; Allen Marr, Geocomp Corporation; Neven Matasovic,
Geosyntec Consultants; Jorge Meneses, Group Delta Consultants, Inc.; Yoshi Moriwaki,
GeoPentech; Adam Perez, Los Angeles Department of Water and Power; Didier Perret, Natural
Resources Canada; Daniel Pradel, University of California, Los Angeles; Michael Riemer,
University of California, Berkeley; Curt Scheyhing, Group Delta Consultants, Inc.; Thomas
Shantz, California Department of Transportation; Sabanayagam Thevanayagam, State University
of New York, Buffalo; Kohji Tokimatsu, Toyko Institute of Technology, Japan; Sjoerd Van
Ballegooy, Tonkin and Taylor; Rick Wentz, Wentz Pacific, LTD; Derek Wittwer, U.S. Bureau of
Reclamation; and Zia Zafir, Kleinfelder, for contributing to meaningful discussions during
meetings and the committee workshop. Many members of the technical community interacted
with members of the committee throughout the conduct of this study. The committee is
appreciative of all this input.

vii
This report has been reviewed in draft form by individuals chosen for their diverse
perspectives and technical expertise. The purpose of this independent review is to provide candid
and critical comments that will assist the institution in making its published report as sound as
possible and to ensure that the report meets institutional standards for objectivity, evidence, and
responsiveness to the study charge. The review comments and draft manuscript remain
confidential to protect the integrity of the deliberative process. We wish to thank the following
individuals for their participation in the review of this report:

Ronald Andrus, Clemson University


Gregory Baecher, NAE, University of Maryland, College Park
Ross Boulanger, University of California, Davis
Jonathan Bray, NAE, University of California, Berkeley
K. Önder Çetin; Middle East Technical University, Turkey
Lloyd Cluff, NAE, Pacific Gas and Electric Company (emeritus)
Misko Cubrinovski, University of Canterbury, New Zealand
Ahmed-Waeil Elgamal, University of California, San Diego
Liam Finn, University of British Columbia, Vancouver (emeritus)
Kenji Ishihara, Chuo University, Tokyo
Michael Lewis, Bechtel Corporation

Although the reviewers listed above provided many constructive comments and
suggestions, they were not asked to endorse the conclusions or recommendations nor did they
see the final draft of the report before its release. The review of this report was overseen by
Robin McGuire, Lettis Consultants International, Inc., and Andrew Whittle, Massachusetts
Institute of Technology, they were responsible for making certain that an independent
examination of this report was carried out in accordance with institutional procedures and that all
review comments were carefully considered. Responsibility for the final content of this report
rests entirely with the authoring committee and the institution.

viii
Contents

SUMMARY .....................................................................................................................................1

1 INTRODUCTION .............................................................................................................14
Liquefaction Hazards, 17
State of Practice for Liquefaction Assessment, 22
Prior Reviews of Liquefaction Assessment, 23
The Committee’s Task, 25
Report Organization, 27

2 A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION ..........................29


Mechanism of Liquefaction, 30
Factors Affecting Liquefaction Potential and Its Consequences, 39
Earthquake Loading, 43
Soil Resistance to Liquefaction, 44
Consequences of Liquefaction, 47

3 CASE HISTORIES ............................................................................................................52


Site Characterization for Case History Assessments, 53
Liquefaction Triggering Case Histories, 58
Site-Specific Ground Motion Recordings, 69
Lateral Spreading Case Histories, 70
Post-Liquefaction Shear Strength Case Histories, 73
Nontraditional Sources of Data, 75
Enhancing Database Development, 75

4 THE SIMPLIFIED STRESS-BASED APPROACH TO TRIGGERING


ASSESSMENT ..................................................................................................................80
Seismic Demand, 81
Liquefaction Resistance, 86
Geotechnical Field Data for Liquefaction Triggering Analyses, 106
Interpreting the Results of Stress-Based Triggering Analysis, 107

5 ALTERNATIVE APPROACHES TO LIQUEFACTION TRIGGERING


ASSESSMENT ................................................................................................................111
Regional Liquefaction Hazard Maps and Past Occurrence of Liquefaction, 112
Cyclic Strain Approach, 113
Energy-Based Approaches, 115
Laboratory and Physical Model Testing, 117
Field Measurement of Pore-Pressure Generation Under Dynamic Loading, 127
Computational Mechanics Approaches toAssess Liquefaction Triggering, 129
ix
6 RESIDUAL SHEAR STRENGTH OF LIQUEFIED SOIL ............................................133
Residual Shear Strength of Sandy Soils, 134
Residual Shear Strength of Gravelly- and Fine-Grained Soils, 140
Dealing with Uncertainty, 141

7 EMPIRICAL AND SEMI-EMPIRICAL METHODS FOR EVALUATING


LIQUEFACTION CONSEQUENCES ............................................................................143
Screening Procedures, 144
Flow Sliding, 149
Lateral Spreading, 150
Liquefaction-Induced Settlement, 158
Deep Foundations, 161
Shallow Foundations, 166
Retaining Structure Damage, 166
Utilities and Buried Structures, 167
Liquefaction-Induced Modification of Ground Motions, 169

8 USE OF COMPUTATIONAL MECHANICS TO PREDICT LIQUEFACTION


AND ITS CONSEQUENCES .........................................................................................170
Computational Liquefaction Modeling in Engineering Practice, 172
Issues in the Computational Modeling of Liquefaction Problems, 174
Constitutive Modeling of Liquefiable Soil, 180
Recent Computational Research Developments Applicable to Liquefaction
Analysis, 182

9 PERFORMANCE-BASED EVALUATION AND DESIGN .........................................189


Approaches to Performance-Based Evaluation, 191
Future Developments for Performance-Based Evaluations, 198

10 RECOMMENDATIONS .................................................................................................200
Collecting, Reporting, and Assessing Data Sufficiency and Quality, 201
Addressing the Spatial Variability and Uncertainty of Data, 205
Improving Tools for Assessing Liquefaction Triggering and Its Consequences, 208
Improving Research and Practice, 210

REFERENCES ............................................................................................................................212

APPENDIXES
A Biographical Sketches of Committee Members ..............................................................248
B Meeting Agendas and Workshop Participants .................................................................254
C Histograms (or parameter distributions) of Recent Liquefaction Triggering Databases .265
D General Description of Performance-Based Design ........................................................271
E Glossary ...........................................................................................................................276

x
Summary

Earthquake-induced soil liquefaction (liquefaction) is a leading cause of earthquake damage


worldwide. Liquefaction is often described in the literature as the phenomena of seismic
generation of excess porewater pressures and consequent softening of granular soils. Many
regions in the United States have been witness to liquefaction and its consequences, not just
those in the west that people associate with earthquake hazards. Earthquakes in 1811 and 1812
caused extensive liquefaction in a region along the Mississippi River, stretching for
approximately 150 km from the Memphis area north (called the New Madrid zone). In 1886 an
earthquake caused widespread liquefaction and associated ground displacements in the
Charleston, South Carolina, area. Liquefaction may have contributed to deaths associated with
tsunamis by destabilizing delta fronts during the 1964 earthquake in Alaska. In 1971, slumping
due to liquefaction nearly resulted in the overtopping of a dam and discharge of the reservoir at
the terminus of the Los Angeles Aqueduct above the San Fernando Valley, an event that would
have flooded thousands of homes in the valley below the dam. The city of Kobe, Japan, has yet
to recover economically from liquefaction-related damage at the city’s port caused by the 1995
Hyogo-ken Nanbu earthquake. Liquefaction during the 2010-2011 earthquake sequence in
Christchurch, New Zealand, led to the loss of 15,000 single-family homes and hundreds of
buildings in the central business district. The most damaging of those earthquakes to the built
environment had a relatively modest magnitude of 6.2.
These examples illustrate the importance of accurate assessments of where liquefaction is
likely and of what the consequences of liquefaction may be. Such assessments are needed to
protect life and safety and to mitigate economic, environmental, and societal impacts of
liquefaction in a cost-effective manner. Assessment methods exist, but methods to assess the
potential for liquefaction triggering are more mature than are those to predict liquefaction
consequences, and the earthquake engineering community wrestles with the differences among
the various assessment methods for both liquefaction triggering and consequences. This report
evaluates these various methods, focusing on those developed within the past 20 years,
and

1
2 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

recommends strategies to minimize uncertainties in the short term and to develop improved
methods to assess liquefaction and its consequences in the long term. This report represents a
first attempt within the geotechnical earthquake engineering community to consider, in such a
manner, the various methods to assess liquefaction consequences.
Liquefaction occurs when stresses and deformation in the ground caused by earthquake
shaking disturb the soil structure (i.e., the arrangement of individual soil grains—the soil fabric)
of saturated, geologically unconsolidated soils. Water in the pore spaces between soil particles
will resist the natural tendency of the soils to consolidate into a denser and more stable
arrangement during shaking. Because the soil cannot change in volume until water is drained
from the pore spaces, porewater pressure will rise, soil particles may lose contact with each
other, and the soil mass may lose much of its strength. This chain of events is referred to as
liquefaction triggering. In this report, the term “liquefaction” refers not only to the conditions at
the onset of liquefaction triggering but also to the phenomena that occur within the soil when in
that triggered state (e.g., flow liquefaction).
When liquefaction triggering occurs, the soil may lose much of its stiffness and strength
and it may also become easier to deform and may flow laterally. Similarly, the soil may also lose
its ability to support an overlying structure or buried utility. Whether or not liquefaction is
triggered largely depends on the characteristics of ground shaking and on the density and initial
effective (interparticle) stress of the soil. Other factors, such as soil type, soil fabric, age, and the
orientation and levels of pre-earthquake shear stresses, also influence the potential for
liquefaction triggering.
Consequences of liquefaction also depend on a host of factors, including soil type and
stratigraphy (layering), ground surface topography, and the engineered infrastructure near where
liquefaction occurs. The ground may be displaced vertically or laterally, landslides may occur,
embankments may slump, foundations may fail, and mixtures of soil and water may erupt at the
ground surface. These effects may lead, in turn, to settlement, distortion, and the collapse of
buildings; the disruption of roadways; the failure of earth-retaining structures; the cracking,
sliding, and overtopping of dams, highway embankments, and other earth structures; the rupture
or severing of sewer, water, fuel, and other lifeline infrastructure; the lateral displacement and
shear failure of piles and pier walls supporting bridges and waterfront structures; and the uplift of
underground structures.
Methods for estimating liquefaction triggering and its consequences vary for many reasons.
Practitioners, clients, and regulatory agencies may differ in their technical or financial capacities.
Projects vary in size, type, and acceptable consequences and risk. What may be acceptable risk
for buildings and roads in sparsely populated regions may not be for dams, bridges, port and
harbor facilities, and nuclear power plants in more populated regions. Project requirements may
differ further with respect to relevant guidance documents, standards, and code provisions.
Analysis of liquefaction and its consequences remains one of the more active areas of
research and development in geotechnical engineering. In 1998 a consensus was reached within
the technical community on the use of an empirical stress-based approach for liquefaction
triggering assessment—called the “simplified method”—first developed in 1971. This method is
still the most commonly used in practice. By 2004, however, reputable groups suggested sets of
changes to this procedure, with many of the groups focused on two alternative approaches. Since
then, both of those alternative approaches have been refined, and additional methods and
modifications have been suggested. Practitioners now must choose among multiple methods
without necessarily understanding their limitations.
SUMMARY 3

The current state of knowledge is limited regarding liquefaction triggering assessment


associated with, for example, the degree of saturation below the groundwater table, at large
depths, beneath sloping ground, in gravelly soils, in soils with a significant component of fine-
grained particles (i.e., silt and clay), in Holocene-age soils due to aging effects, or with
Pleistocene-aged (and older) soils. Few data are available to validate procedures for predicting
the consequences of liquefaction triggering. Large uncertainties associated with the application
of methods to predict both liquefaction triggering and its consequences exist.

THE STUDY CHARGE

In 2010 an ad hoc committee of the Earthquake Engineering Research Institute (EERI)


called for a community-based consideration of the state of the art in liquefaction assessment to
be conducted under the auspices of the National Academies of Sciences, Engineering, and
Medicine (the National Academies). Such a study would help the technical community rebuild
consensus on issues related to liquefaction triggering assessment and foster confidence in
methods used to assess liquefaction triggering and its consequences. Under the sponsorship of
the U.S. Bureau of Reclamation, Nuclear Regulatory Commission, Federal Highway
Administration, American Society of Civil Engineers, Port of Long Beach (California), Port of
Los Angeles (California), and Los Angeles Department of Water and Power, the National
Academies convened a committee of 12 engineers and scientists to solicit input from the
technical community and to critically examine the state of practice in liquefaction triggering and
consequence assessment (see Chapter 1, Box 1.4 for the full statement of task). The committee
was tasked to evaluate the following: the sufficiency, quality, and uncertainties associated with
laboratory and in situ field tests, case history data, and physical model tests for understanding
liquefaction triggering and post-triggering soil behavior; methods to analyze the data from those
tests; and the adequacy and accuracy of empirical and mechanistic methods to evaluate
triggering and resulting deformations in the soil and the structures built in, on, and of those soils.
The committee’s report focuses on developments since the 1996 and 1998 National Science
Foundation/National Center for Earthquake Engineering Research (NCEER) workshops where
consensus was last reached on the topic of assessing liquefaction triggering. The report considers
future directions for research and practice related to collecting, reporting, and assessing the
sufficiency and quality of different types of data, to assessing the spatial variability and
uncertainty in the data, and to developing more accurate assessment tools. The purpose of the
study was not to adjudicate disputes regarding any specific methods. The committee’s report
addresses liquefaction of only naturally occurring soils and fills composed of such soils (i.e., the
report does not extend to mine tailings, solid waste, or stabilized soils). Discussion of remedial
measures to increase liquefaction resistance and to mitigate post-liquefaction consequences is
also outside the study scope. Recommendations in the report focus on how engineering practice
with respect to liquefaction triggering and consequence assessment could be improved using
existing tools and on future research that could improve practice.
4 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

KEY FINDINGS AND CONCLUSIONS

Key findings and conclusions from each chapter of the committee’s report are listed in the
beginning of each chapter of the report and summarized in this section. These findings and
conclusions, discussed in greater detail in the report, support the committee’s final
recommendations presented in this summary and in Chapter 10.
To evaluate liquefaction hazards, it is necessary to assess the susceptibility of a soil to
liquefaction, the potential for liquefaction triggering, and the potential consequences of
triggering given anticipated ground motions at a site. Not all soils are susceptible to liquefaction,
but granular soils such as sands, silty sands, nonplastic silts, and gravels confined by lower-
permeability layers are known to have liquefied in past earthquakes. Saturation is considered to
be a necessary precondition for liquefaction. Laboratory testing currently cannot be used as the
primary means to assess liquefaction potential in truly cohesionless soils. Instead, in situ testing
of soil to assess its resistance to liquefaction is commonly used (e.g., through standard
penetration testing [SPT], cone penetration testing [CPT], or shear wave velocity [Vs] testing), in
spite of their own limitations.
The basis of the most commonly used methods to predict liquefaction triggering in
engineering practice is a simplified stress-based approach developed over 40 years ago. Whereas
this approach is useful, its results are fraught with uncertainty. Methods used to predict the
consequences of liquefaction triggering also include simplified approaches with high degrees of
uncertainty. Additionally, the methods commonly used to predict liquefaction triggering and its
consequences are generally not compatible with the current trend toward risk-based evaluation of
seismic hazard (sometimes referred to as performance-based seismic design when referring to
engineering design that is required to meet a certain measurable performance criteria).

Uncertainties in Assessments

Uncertainties are introduced into liquefaction assessments from many sources, beginning
with unknowns related to site geology (e.g., the vertical and lateral extents of specific soil layers
and groundwater conditions). The in situ test methods used to determine soil liquefaction
resistance introduce additional uncertainties given the lack of truly standardized protocols for
their use, as well as the lack of protocols to characterize relevant soil and profile properties (e.g.,
anisotropy in soil fabric; structure of gravelly soils). Incomplete knowledge of the influence of
earthquake magnitude or duration, ground motion intensity, the in situ effective overburden
stress, non-level ground conditions, and the amount of fine-grained particles in the soil introduce
more uncertainties.
Case histories of liquefaction are important in developing and validating liquefaction
analysis procedures. Lack of protocols for documenting and rating the quality of case histories,
however, introduce uncertainties. The same case history may be documented differently in
databases assembled by different investigators. Even with a consistent standard for case history
documentation, inconsistent information and interpretations can be found among the databases.
For example, whether or not and where liquefaction has occurred or the location at which it has
occurred within a soil profile; weathering and cementation characteristics; estimates of
overburden stress; the layers most susceptible to liquefaction and their properties; the
identification of geologic controls on soil properties that might affect liquefaction (e.g., lateral
SUMMARY 5

variation in the shape and soil content of a layer); the age of a soil; and estimates of prior seismic
shaking have all been inconsistently interpreted or documented. These inconsistencies
will contribute to uncertainties in liquefaction relationships developed from the case histories.
Confidence in relationships derived from case history data (e.g., deterministic plots that
relate resistance of soil to liquefaction and in situ test indices) is also affected by the lack
of high-quality case history data points that plot close to and on either side of boundary lines
that separate liquefaction and no-liquefaction zones. In particular, databases need case
histories for critical depths (i.e., those at which liquefaction is known to have occurred) greater
than about 15 meters, for earthquake magnitudes less than approximately 5.9 and greater than
approximately 7.8, and for cases where liquefaction may reasonably have been expected but was
not observed.

The Simplified Stress-Based Approach for Assessing Liquefaction Triggering

Several empirical liquefaction triggering relationships that are derived primarily from the
case history data are applied to liquefaction assessments conducted with the simplified stress-
based approach. Assumptions inherent with use of the simplified stress-based approach include a
horizontally layered soil profile, horizontal ground surface, vertically propagating shear waves,
and undrained soil response. These relationships differ in the use of the in situ
parameter correlated to the soil’s cyclic resistance ratio (CRR), the various adjustment factors
incorporated in the calculation of the CRR, and the earthquake-induced cyclic stress ratio
(CSR) of the soil. Because each relationship was developed using an associated suite of
adjustment factors, using adjustment factors developed for one relationship may result in
increased uncertainty and error if applied to another relationship. Going forward, only the
specific adjustment factors used for development of a particular liquefaction triggering
relationship should be used with that relationship.
Some adjustment factors are not well constrained over the entire range of engineering
interest by the empirical data (e.g., the stress magnitude adjustment factor, K). In cases
where factors are not well constrained by empirical data, adjustment factors should be developed
using experimental data (including centrifuge and shaking table experiments) and
engineering mechanics principles; they should not simply be based on statistical extrapolation of
the data. No consensus exists regarding the appropriate adjustment factors for the presence of an
initial static shear stress (i.e., K). Additional data and research are needed to allow better
understanding of these effects. Other issues that require additional research include
adjustment factors for the magnitude of the normal stress, the influence of geologic age on
liquefaction resistance, the influence of earthquake magnitudes less than 7.0 or greater than 8.0,
and how best to estimate the liquefaction potential of gravelly soils and soils with a
substantial amount of silt and clay particles. Unbiased model parameters need to be
employed in developing CRR curves and adjustment factors. The use of biased
estimates of parameters (e.g., of the depth adjustment factor, rd) serves only to increase
uncertainty regarding the results of an analysis. Significant reduction in uncertainties
associated with the various adjustment factors used in the simplified approach is most
dependent on expansion and improvement of the case history database, but mechanics-
based methods can be used where data are insufficient.
6 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Alternative Approaches to Liquefaction Assessment

Methods other than the simplified stress-based approach to assess liquefaction triggering
include a cyclic strain-based alternative to the cyclic stress-based approach, energy-based
approaches, laboratory tests and physical models, and computational mechanics-based methods.
Further development of these methods could improve the accuracy and reliability of liquefaction
assessment. Other tools, such as regional maps that relate liquefaction hazards to geologic map
units, are used as screening and planning tools and in regional damage and loss assessments.
Maps of probabilistic liquefaction hazards could be expanded in coverage, even to a national
scale similar to the national seismic hazard maps.
The complexity of phenomena subsequent to liquefaction triggering has prompted the
development of various empirical and semiempirical procedures for use by earthquake engineers
to assess potential consequences of liquefaction. These include screening procedures that employ
damage indices to determine if the consequences of liquefaction are of engineering concern;
procedures to evaluate the potential for lateral spreading, flow sliding, and slope instability; and
procedures to evaluate the performance of deep and shallow foundations, earth-retaining
structures, subsurface utilities, and buried structures. While most of these procedures are
logically formulated, and some are supported by laboratory and model testing, they have
undergone little field validation. The limitations of simple criteria and procedures to screen sites
for severity and damage potential from liquefaction triggering are not often stated explicitly, and
the applicability of many of those procedures has not been rigorously addressed. Similarly,
empirical and semiempirical models to evaluate such liquefaction impacts as flow sliding, lateral
spreading displacement, loss of bearing capacity, and increased lateral earth pressures and
buoyancy effects, and to model loads induced on structures by liquefied soils, generally have not
been validated rigorously. Lack of rigorous validation is due largely to the paucity of
appropriately documented case histories of liquefaction consequences.

Computational Methods

Empirical and semiempirical approaches cannot account for all of the effects attributable to
site-specific geology and topography, engineered structure configurations, and ground motion
characteristics. When combined with appropriate generalizations of fluid flow relative to the
solid phase, computational models can be used to solve boundary-value problems such as the
deformation and pore-pressure response of a soil deposit with an overlying structure subjected to
earthquake loading. Computational methods that use the principles of mechanics and incorporate
appropriate constitutive relations (i.e., mathematical equations that describe the response of the
soil to an external stress or strain) applied to a properly characterized site (i.e., a site
characterized to a level of detail appropriate to the level of analysis) provide a means to describe,
in detail, pore-pressure generation, redistribution, and dissipation; void (porosity) redistribution;
soil and porewater flow; and soil-structure interactions. The accuracy with which a
computational method describes the response of the soil to external loads (e.g., to the
earthquake-induced ground motion) depends largely on the accuracy of the constitutive relations
and input parameters employed.
Continued development of new, more detailed, and more accurate constitutive and
computational models for assessing liquefaction triggering and consequences are needed. In
SUMMARY 7

particular, constitutive models are needed that describe the behavior of soils near and after the
triggering of liquefaction, as are computational models that capture post-triggering soil behavior.
These computational models need to be supplemented by laboratory and physical model testing
that both elucidate key aspects of the behavior of soils near and subsequent to liquefaction
triggering and can be used to validate the constitutive and computational models.

Performance-Based Evaluation and Design

Current application of the stress-based method for liquefaction triggering assessment


focuses on computing deterministic factors of safety versus depth for a given earthquake
scenario. To be compatible with current developments in other areas of earthquake engineering
(e.g., performance-based design of structures), the practice of liquefaction assessment needs to
move toward fully probabilistic analyses that incorporate the complete range of possible
damaging earthquake ground motions (in terms of both ground motion intensity and earthquake
magnitude), their probable frequency of occurrence, and the variability in the parameters and
adjustment factors used to estimate the CRR. These probabilistic analyses can incorporate the
epistemic uncertainty among the available empirical models by using a logic tree approach that
can also be used to consider uncertainty in the site characterization. The uncertainties involved in
the assessment of earthquake ground motions, system response, physical damage, and losses
make probabilistic methods for liquefaction consequence assessment central to performance-
based evaluation and design. Until fully probabilistic liquefaction hazard analyses (PLHAs) are
developed, deaggregation data can be used in conjunction with probabilistic seismic hazard
analysis results to approximate the liquefaction hazard.
Fully probabilistic procedures (i.e., PLHAs), in which contributions from all potential levels
of ground motion are considered, have been developed and shown to produce more complete and
rational estimates of liquefaction hazards in different seismic environments than do deterministic
methods. PLHAs are well developed for liquefaction triggering and recently have been extended
to consequences such as lateral spreading and post-earthquake free-field settlement. They have
also been used to show that estimates of actual liquefaction hazards using current procedures as
applied in practice can be highly inconsistent. Consistent application of conventional procedures
can result in excessive conservatism in some seismic environments and excessive hazard and risk
in others; under such conditions, available funds for seismic retrofit programs, for example,
cannot be allocated efficiently.
Advances in performance-based procedures for liquefaction problems will require improved
understanding of the potential damage to different types of structures and facilities given
different levels of liquefaction-induced ground deformation and improved understanding of the
costs and time requirements to repair liquefaction-associated damage. Performance-based
approaches offer great opportunities to produce safer, more resilient structures and facilities and
to more efficiently use available resources. Adoption of performance-based engineering concepts
will require major changes in the thinking, practice, and education of engineers. It represents
moving away from dependence on rules of thumb and factors of safety and toward a design
process rooted in the probabilistic prediction of the behavior of engineered systems and
estimation of economic losses. The adoption of performance-based liquefaction hazard
evaluation in practice, however, will require computing capacity sufficient to perform the
voluminous calculations involved in probabilistic performance-based frameworks.
8 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

RECOMMENDATIONS

Findings and recommendations provided in this report are based on the study committee’s
assessment of past and current research. As such, recommendations in this report reinforce old
ideas or are intended to expand or provide new directions for current research and practice.
Given the nature and complexity of liquefaction phenomena and their potential impacts on life
safety and infrastructure, improvement of liquefaction assessment methods will require
multidisciplinary effort among engineers, geoscientists, infrastructure owners, and major
stakeholders. Many efforts will require strong community-based collaborations that enable
holistic understanding of problems associated with liquefaction. Nevertheless, it is beyond the
scope of the committee’s statement of task to make recommendations regarding the levels and
organization of funding necessary to implement these recommendations.

Recommendation 1. Establish curated, publicly accessible databases of relevant


liquefaction triggering and consequence case history data. Include case histories in which
soils interact with built structures. Document the case histories with relevant field,
laboratory, and physical model data. Develop the databases with strict protocols and
include indicators of data quality.

Field data from liquefaction case histories are important for the development, calibration,
and validation of the various approaches to predicting liquefaction triggering and its
consequences. Case histories are needed with parameter values beyond the ranges for which
sufficient data are currently available. Incorporating into databases data collected since 1995
would enhance the value of the databases. More case histories are needed that represent
behaviors of liquefiable soils at depths greater than approximately 15 meters; for soils subjected
to relatively small and large magnitude events; of soils with greater than 35% fines content; and
of soils with a greater than 50% fine-grained particles of low plasticity. More case histories
representing soils in sloping ground or adjacent to free faces in potentially liquefiable terrain
with SPT blow count values of greater than 15 blows per 30 cm or CPT tip resistance values of
greater than 85 atmospheres (8.6 MPa) are also needed. Including additional case histories where
there are no surface manifestations of liquefaction or its consequences will also add to the body
of knowledge by increasing an understanding of where and why liquefaction does not occur.
Furthermore, there is need for greater consistency of the data included in the databases, more
transparency related to data quality, and greater ease of access to the databases by researchers
and practitioners.

Recommendation 2. Characterize locations with high probability of liquefaction and


establish them as field observatories for liquefaction triggering and its consequences.

Liquefaction case histories are usually developed after an area has been subject to ground
motion and, in most cases, after observation of surface manifestations of liquefaction. In such
cases, pre-liquefaction soil properties, changes in the soil properties, and ground displacements
can only be inferred. Establishing observatories at well-characterized sites with a high
probability of liquefaction-inducing earthquakes can result in better knowledge about pre-
liquefaction soil properties, and fill gaps in case history databases. High-quality site
characterization of surface and subsurface conditions would be conducted prior to or as part of
SUMMARY 9

establishment of the observatory. Post-event data from these sites could be used to calibrate and
validate procedures for assessment of liquefaction triggering and its consequences and for
providing needed information regarding liquefaction behavior at depth, in gravels and fine-
grained soils, on sloping ground, and within and beneath embankments, as well as information
regarding the soil-structure interaction effects in liquefied soil.

Recommendation 3. Use data from the cone penetration test (CPT) for field-based
estimates of liquefaction resistance where feasible. If the standard penetration test (SPT) is
used for this purpose, make hammer energy measurements. Supplement field-based
estimates with other methods as appropriate to characterize the site.

CPT soundings offer advantages over other methods of estimating liquefaction resistance in
both the detection of thin layers that may affect liquefaction triggering and subsequent pore-
pressure redistribution and in the reproducibility of results. CPT results are less dependent on the
equipment operator or setup than most other in situ test methods, and CPT can be performed
with relative speed and economy. It may also be prudent to conduct one or more SPT borings in
addition to CPT soundings to provide physical samples for soil characterization as well as to
provide an additional means of liquefaction potential assessment. If SPT measurements are
performed, hammer energy measures should be made and the SPT setup should minimize the
need for additional correction factors. The CPT can also provide a cost-effective means to
measure Vs that can enhance site characterization and liquefaction assessment. Nevertheless,
CPT soundings are not feasible in very dense or gravelly soils. In such cases, other methods need
to be used to estimate liquefaction resistance. In gravelly soils, instrumented Becker hammer
measurements can serve the same purpose as the SPT, and it may be the best available method
for soils that cannot be penetrated by the CPT. The use of multiple independent methods can
provide additional data to inform the liquefaction potential assessment.

Recommendation 4. When refining or developing new empirical relationships for use in


liquefaction analyses, incorporate unbiased estimates for input parameters; identify and
quantify when possible the uncertainty associated with those estimates; and use soil
mechanics principles, seismologic principles, and experimental data to extrapolate beyond
ranges in which field data constrain the empirical relationships.

Identifying and, if possible, quantifying the uncertainty associated with each adjustment
factor, empirical correlation, and parameter relationship used in a liquefaction assessment will
help to avoid the compounding of uncertainties and to facilitate assessment of the overall
uncertainty associated with the method. When developing a liquefaction triggering or
consequence assessment method, using adjustment factors, empirical correlations, and parameter
relationships with built-in bias will compound uncertainties and make it difficult, if not
impossible, to assess the overall uncertainty associated with the method—an important
consideration in liquefaction assessments.

Recommendation 5. Use geology to improve the geotechnical understanding of case


histories and project sites, particularly where potentially liquefiable soils vary in thickness,
continuity, and engineering properties.
10 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

The potential for liquefaction and consequent ground failures can depend greatly on
geologic variability in the thickness, lateral extent, and continuity of liquefiable soils.
Sedimentary stratigraphy is related to depositional setting. For example, understanding the shape
and extent of a thin sand body deposited in the channel of a meandering river may provide
information about the potential for lateral spreading. Geology also affects liquefaction potential
through weathering and consolidation, resulting in an increase in the resistance of a soil to
liquefaction with age. Understanding geology at and below the ground surface feeds into
regional liquefaction hazard maps that relate surficial geology to probabilistic estimates of
liquefaction potential. The geological traces of liquefaction can aid in evaluating liquefaction
potential where written history contains few if any earthquakes. A database of such
paleoliquefaction evidence has been compiled for the central and eastern United States. This
database could be expanded nationwide or beyond, both to expand the temporal range of
liquefaction case histories and to use sedimentary environment, depositional age, and tectonic
setting as screening criteria in assessments of liquefaction hazards.

Recommendation 6. Implement simplified stress-based methods for liquefaction triggering


in a manner consistent with how they were developed. Avoid using techniques and
adjustment factors from one variant of a method with other variants. Consider using more
than one simplified method when making a liquefaction triggering assessment.

Significant epistemic (modeling) uncertainties are associated with all variants of the
simplified stress-based approach to liquefaction triggering assessment. Combining techniques or
factors from different variants may serve to compound the uncertainties in unknown and
unquantifiable ways. Use of multiple variants of the simplified method may be warranted if
analysis results from one variant indicate a marginal factor of safety, where consequences of
liquefaction are considered unacceptable, or when a method is applied beyond the bounds where
method relations are constrained by the data. A more sophisticated liquefaction assessment may
be warranted if results from use of multiple variants are contradictory.

Recommendation 7. In developing methods to evaluate liquefaction triggering and its


consequences, explicitly incorporate uncertainties from field investigations, laboratory
testing, numerical modeling, and the impact of the local site conditions on the earthquake
ground motions.

Descriptions of procedures to evaluate liquefaction do not always state clearly where and
what the uncertainties are and typically do not address how to incorporate them into the
liquefaction assessment. Uncertainty in correction factors (e.g., corrections applied to SPT blow
counts, magnitude scaling factors) is rarely accounted for, and in some cases, it may introduce
even more uncertainty in the results of the analysis than that associated with the original
measurement. Uncertainty in a liquefaction consequence or triggering assessment increases
rapidly as the assessment method moves beyond the range within which it is constrained by field
data. Uncertainties in field data and analytical and experimental results need to be stated in the
form of error bounds, standard deviations, bias, or other statistically appropriate measures, and
liquefaction assessment procedures need to clearly identify the range over which they are
constrained by both field and laboratory test data. Users, in turn, need either to refrain from
SUMMARY 11

extrapolating beyond this range or to explicitly identify when they have done so, and they need
to qualify the results with, for example, laboratory test results or other information.

Recommendation 8. Refine, develop, and implement performance-based approaches to


evaluating liquefaction, including triggering, the geotechnical consequence of triggering,
structural damage, and economic loss models to facilitate performance-based evaluation
and design.

Regional variations in seismicity, coupled with the significant uncertainties in earthquake


loading and liquefaction resistance, warrant probabilistic characterization for the evaluation of
both the hazards of liquefaction triggering and the risk of liquefaction triggering. The
geotechnical community needs to develop PLHAs to meet the growing demand from the greater
earthquake engineering community for risk-based liquefaction assessment as part of the trend
toward performance-based design. The results of these analyses should be expressed in the form
of a response curve that indicates how often different levels of response (e.g., liquefaction
triggering, lateral spreading displacement, settlement) can be expected to occur. In a
performance-based framework, the results of this type of analysis can be convolved with damage
and loss models to estimate the risk associated with liquefaction. Risk can be expressed in terms
of a risk curve that indicates how often different levels of loss can be expected to occur. Losses
include loss of life, loss of functionality, and direct and indirect economic consequences.

Recommendation 9. Use experimental data and fundamental principles of seismology,


geology, geotechnical engineering, and engineering mechanics to develop new analytical
techniques, screening tools, and models to assess liquefaction triggering and post-
liquefaction consequences.

New approaches to liquefaction assessment are needed that go beyond the empirical
approaches commonly used in practice. These new approaches need to be based on sound
science (e.g., geology, seismology, and physics) and on fundamental principles of engineering
mechanics and geotechnical engineering. Such approaches need to: quantitatively relate
liquefaction hazards to geologic units at appropriate depths; increase predictive capabilities in
evaluations of both triggering and its consequences; be compatible with probabilistic
characterization of seismic loading; be consistent with performance-based assessments of
liquefaction consequences; and be readily accessible to practicing engineers. Development and
implementation will require greater collaboration among geotechnical engineers and engineering
geologists to characterize the depth-dependent distribution and properties of geologic units. The
approaches need to be consistent with patterns of behavior identified through analysis of field
case histories and laboratory and physical model test data. Strain-based and energy-based
approaches are two avenues of development that merit consideration.
Rigorous validation of these models, using case history data and development of
implementation approaches that make the models accessible to practitioners, is needed to
facilitate their adoption in engineering practice. Use of such integrated and validated models in
practice would be consistent with trends in other areas of modern earthquake engineering
practice.
12 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Recommendation 10. Develop and validate computational models for liquefaction analyses.
Use laboratory and physical model tests at different spatial scales and case histories to
provide insight into fundamental soil behavior and to validate the application of
constitutive models to boundary-value problems.

Computational models are valuable tools when empirical models give contradictory or
inconclusive results, or when a more detailed assessment of site and structure response is needed.
However, computational models do not always accurately reproduce behavior across laboratory-
test, physical-test, and field scales. They are not always able to capture complex material
behavior under cyclic loading, including the intimately coupled deformation and flow patterns
before and after triggering. Although models based on discrete element mechanics offer a
promising avenue for modeling mechanisms at the grain level, at present they are
computationally burdensome. Further developments are needed to enhance the accuracy and
computational efficiency of discrete methods and to couple the discrete methods with
simulations across scales. Meshless methods offer the potential to predict flow behavior of soil
subsequent to liquefaction. To realize this potential, however, further insight from grain-scale
models (e.g., through the use of multiscale methods) is needed. Rigorous validation procedures
that distinguish between the accuracy of the computational method and the complexity of the
problems for which the predictive model has been validated are required for all computational
procedures, regardless of the computational approach and method of analysis used.

Recommendation 11. Conduct fundamental research on the stress, strain, and strength
behaviors of soils prior to and after liquefaction triggering; devise new laboratory and
physical model experimental techniques to aid development of constitutive models of those
behaviors.

Advancing the state of the art and practice in liquefaction assessment requires fundamental
research on liquefaction phenomena, both theoretical and experimental, because the stress-strain
behavior of soil prior to and following liquefaction triggering is complex and not fully
understood. No available constitutive model captures all of the relevant soil behavior within a
rigorous physics-based framework. Research is needed on soil dilation prior to and after
liquefaction triggering, on the solid-to-viscous-liquid behavior transformation that can
accompany triggering, and on the mechanisms that control the post-triggering fabric degradation
that influences stress-strain behavior of soils. Mechanisms of post-triggering particle
rearrangement and reconsolidation, as well as the effects of relative density and consolidation
stress on these mechanisms, are not well understood and are in particular need of further
research.

IMPROVING FUTURE RESEARCH AND PRACTICE

Liquefaction assessment will be improved if informed by more and better data, by


quantification of uncertainties, and by fundamental scientific and engineering principles.
Engineering practice will advance when new methods are developed that take into account the
needs and resources of practicing engineers and are designed to be easily implementable. The
fundamental basis and applicability of new assessments methods need to be understandable, and
SUMMARY 13

outputs of models need to be translatable to problems of interest. Limitations of models need to


be clearly defined so that model reliability can be understood under given circumstances.
Liquefaction assessment approaches also need to keep pace with other aspects of earthquake
engineering practice—for example, by being compatible with performance-based design.
Advancing the state of knowledge and practice related to liquefaction assessment will
require concerted efforts not only by researchers and engineering practitioners but also by
facility owners and stakeholders at large. Researchers can provide more high-quality field case
history data supplemented with laboratory and model test data as well as more accurate models
to assess earthquake-induced soil liquefaction. Practitioners need to better understand the
limitations of characterization and assessment approaches and the geologic and tectonic controls
on liquefaction hazards at their sites. Facility owners can support the establishment of public
case history databases and of well-characterized and instrumented field observatories for the
study of liquefaction effects. Stakeholders such as regional planners could support the
development of region-wide liquefaction hazard mapping. It is only through collaborative and
interdisciplinary efforts that liquefaction triggering and consequence assessment can undergo the
fundamental changes necessary to substantially improve practice.
1
Introduction

An earthquake on the morning of February 9, 1971, produced a landslide in the upstream


face of Lower San Fernando Dam (also known as the Lower Van Norman Dam), an earthen
structure located at the terminus of the Los Angeles Aqueduct (see Figure 1.1). The dam crest
dropped approximately 10 meters, leaving only about 1.5 meters of broken soil as freeboard to
prevent overtopping and release of a reservoir holding approximately 15 million tons
(approximately 13.6 million metric tons, or 4 billion gallons) of water (see Figure 1.1). Eighty
thousand people were evacuated from the valley below. Many lives could have been lost had the
water level in the reservoir been at maximum allowable pool level (Cortright, 1975) or the slide a
bit larger (Page et al., undated). It would have been among the greatest earthquake-related
disasters in U.S. history.
This near-disaster resulted from earthquake-induced liquefaction. Seismic shaking of
sufficient strength and duration can transform certain saturated but firm soils1 into a suspension
of soil particles and water that behaves in a manner similar to a viscous fluid. Under these
conditions, the once-firm soil is transformed to one with low strength and stiffness. This
phenomenon is called liquefaction (see Box 1.1), and in this report, the term “liquefaction”
encompasses not only the conditions at the onset of phenomena (i.e., liquefaction triggering) but
also the associated soil behaviors while in that state (e.g., flow liquefaction). At the Lower San
Fernando Dam, the earthquake caused the soil within the dam to liquefy, and the loss of soil
strength set off the landslide. The dam crest began to lower shortly after earthquake shaking
began (Seed et al., 1975a; Castro et al., 1985; Gu et al., 1993). This example underscores the
importance of assessing liquefaction hazards, both by determining whether a soil is likely to
liquefy and by estimating consequences that may include, for example, damage to an earthen
dam.

1
Soils, in this report, refer to particulate mineral Earth materials less consolidated than rock, as per geotechnical
engineering practice. Soils contain readily visible particles (e.g., gravel and sand), particles visible only with
magnification (e.g., silt and clay), or mixtures thereof.

14
INTRODUCTION 15

Dam
(a) (b)

FIGURE 1.1 The Lower San Fernando Dam in Los Angeles (a) prior to the 1971 San Fernando
earthquake when water levels were high, showing downstream residential development. Water levels
shown here were about 10 meters lower than those at the time of the earthquake. SOURCE: USGS
(b) the dam showing damage resulting from the earthquake. Water levels were lowered further following
the earthquake-induced damage. Earthquake-induced liquefaction resulted in the dam crest dropping 10
meters, leaving only 1.5 meters of freeboard as protection against flooding the downstream region with its
population of approximately 80,000 people. SOURCE: Nisee/Peer, University of California, Berkeley.

Accurate liquefaction assessment is essential to protect life and safety, to reduce seismic risk
in a cost-effective manner, and to guide post-earthquake recovery. It is important for new
construction and for enhancing the seismic safety of existing structures. However, liquefaction
assessment can be challenging. The factors that govern liquefaction are often difficult to measure
or predict, and the results of liquefaction analyses in engineering practice are sometimes
contradictory: while one approach might indicate low liquefaction potential, another might
warrant ground improvement costing millions of dollars. This report has its origins in such
dilemmas that combine risk and uncertainty.
16 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

BOX 1.1
What Is Earthquake-Induced Liquefaction?

Earthquake-induced liquefaction refers to the phenomena of seismic generation of excess porewater


pressures and consequent softening of saturated granular soils. The material is typically sand, and less
commonly silt or gravel. Liquefaction results from the tendency of the granular structure of a saturated soil
to collapse and contract, or densify, when subjected to cyclic loading (i.e., the application of repeated
shear stresses) caused by earthquake shaking. Prior to liquefaction, the contact forces between soil
grains provide the stiffness and strength of a saturated soil (Figure 1. The tendency for contraction of the
soil structure (or soil skeleton) leads to a transfer of the load that had been carried by the granular
structure to the porewater filling the voids between individual soil grains, resulting in an increase in the
porewater pressure. When the porewater pressure rises, the contact forces between the grains are
reduced, the soil is more easily deformed, and, in the limiting case, the soil particles may lose contact
completely with each other and go into suspension (Figure 2). The looser the soil, the greater the
potential for contraction under cyclic loading, and the easier it is for the soil to liquefy.
The loss of soil strength associated with liquefaction can lead to large ground deformations and the
inability of the liquefied soil to support overlying materials, including man-made structures. Collateral
consequences can include loss of life and destruction of homes and critical infrastructure. Recovery
following earthquake-induced soil liquefaction can be expensive (tens of millions to billions of dollars) and
may take many years.
The figure offers a simplistic description of the liquefaction phenomenon and should not be
interpreted to suggest that the density of the soil always increases after a soil liquefies. Upward flow due
to porewater pressures generated by the earthquake can loosen soils such that even after resettlement,
soil density is less than it was initially. In some cases, lateral flow after a soil liquefies may also result in a
reduction in soil density due to a phenomenon called dilation (see Chapter 2), wherein a soil expands in
volume when loaded in shear.

(1) (2) (3)


FIGURE Idealized schematic of liquefiable soils. (1) Before an earthquake, individual soil grains are held
in place by frictional or adhesive contact forces, creating a solid soil structure with water filling the spaces
between the grains. Note the grain-to-grain contact. (2) After initiation of liquefaction, particle
rearrangement with no change in volume (e.g., a lateral shift of a half diameter of every other row of
particles in the figure) causes the particles to lose contact and go into suspension, causing the porewater
pressure to increase as gravity load is transferred from the soil skeleton to the porewater. (3) As water
flows out of the soil, the soil particles settle into a denser configuration, the soil skeleton once again
carries the load, and the porewater pressure decreases (back to the initial steady state value).
INTRODUCTION 17

LIQUEFACTION HAZARDS

Large regions of the conterminous United States are susceptible to earthquake shaking strong
enough to cause liquefaction in water-saturated soils. The shaded areas in Figure 1.2 highlight
parts of the lower 48 states where the horizontal component of the peak ground acceleration2
(PGA) specified by the International Building Code (ICC, 2011) for residential and commercial
construction is equal to or greater than 0.1 g (where g is the acceleration due to Earth’s gravity).
In those shaded areas, liquefaction may be of concern wherever potentially liquefiable soils are
present. (Note, however, depending on site-specific conditions, liquefaction cannot be precluded
in the areas that are not shaded.) Historical and geologic evidence attest to the occurrence of
liquefaction in about one-third of the lower 48 states. Table 1.1 provides representative examples
of where liquefaction has occurred prior to 1964. These examples were identified through
historical accounts or paleo-liquefaction investigations. The table includes cases in Arkansas,
California, Illinois, Indiana, Kentucky, Louisiana, Massachusetts, Mississippi, Missouri, New
Hampshire, Oregon, South Carolina, Tennessee, Virginia, and Washington (see Table 1.1).
The consequences of liquefaction include lateral and vertical ground displacements,
landslides, slumping of embankments, foundation bearing failures, ejection of soil-water
mixtures at the ground surface, buoyant uplift of buried structures, and increased lateral earth
pressures on walls. These, in turn, can lead to settlement, distortion, and lateral displacement of
the ground and the collapse of buildings; failure of earth-retaining structures; cracking, sliding,
and overtopping of dams and highway embankments; rupture or severing of sewer, water, fuel,
and other lifeline infrastructure; lateral displacement and shear failure of piles and pier walls
supporting bridges and waterfront structures; and uplift of below-grade structures (see Figure 1.3

FIGURE 1.2 Map showing areas where there is potential for ground shaking intense enough to cause
liquefaction in water-saturated granular materials. The shaded areas indicate where, on soft rock, there is
a 1-in-50 chance in 50 years that peak ground acceleration will attain or exceed one-tenth the
acceleration due to gravity (modified from Petersen et al., 2014). Alaska and Hawaii are not shown, but
those states also have extensive areas of hazard. Note that liquefaction cannot be precluded in areas
where not shaded, depending on site specific conditions. SOURCE: USGS

2
Peak ground acceleration (PGA) is a most common measure of the intensity of ground shaking in a given area. The
PGA is often used in engineering applications, e.g., liquefaction analyses, and in building codes.
18 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

TABLE 1.1 Example Evidence for Pre-1964 Liquefaction in the Conterminous United States
State(s) Evidence References
Arkansas, Sand boils from the 1811-1812 New Reviewed by Tuttle and Hartleb
a
Kentucky, Madrid earthquake and from prehistoric (2012)
Mississippi, predecessors
Missouri, and
Tennessee

California Sand intrusions ascribed to prehistoric Lienkaemper and Williams


earthquakes of the past 1,800 years along (2007)
the Hayward fault

Sand boils as much as 1,500 years old Sieh (1978)


interpreted as evidence for earthquakes
along the San Andreas fault north of Los
Angeles

Ground failures during the 1872 Owens Amos et al. (2013)


Valley earthquake

Ground failures during the 1906 San Youd and Hoose (1978)
Francisco earthquake

Ground failures during the 1933 Long Southern California Earthquake


b
Beach earthquake Center

Illinois and Prehistoric sand dikes 1,500–7,500 years Obermeier et al. (1991)
Indiana old in the Wabash Valley

Massachusetts Historical accounts of ground failures Tuttle and Seeber (1991)


and New during the 1727 Newbury earthquake and
Hampshire sand dikes interpreted as evidence for
preceding earthquakes in the past 4,000
years

Oregon and Sand intrusions along the Columbia River, Obermeier and Dickenson
Washington ascribed to the 1700 Cascadia (2000); Takada and Atwater
earthquake (2004)

South Carolina Ice-age lakebeds repeatedly deformed Janecke et al. (2013)


17,000–23,000 years ago

Sand boils from the 1886 Charleston Dutton (1889); Talwani and
earthquake and from preceding Schaeffer (2001)
earthquakes of the past 6,000 years

Washington Sand boils near Puget Sound, dated to Martin and Bourgeois (2012)
about 1,100 years ago

Ground failures from the Puget Sound Chleborad and Schuster (1998)
earthquakes of 1949 and 1965
a
This comprehensive report also reviews and catalogs evidence for liquefaction in the Wabash Valley and Newbury
areas and additional parts of eastern North America: Charlevoix Seismic Zone of Quebec Province; Central Virginia
Seismic Zone; Marianna, Arkansas; St. Louis, Missouri; and an area in southern Arkansas and northern Louisiana.
b
See http://www.scec.org/education/080307longbeach.html.
INTRODUCTION 19

for examples). In an underwater setting, liquefaction-induced landslides can cause fast-arriving


tsunamis on nearby shores and can be transformed into turbidity currents downslope that break
submarine cables (Seed et al., 1988; Finn, 2003; Tinti et al., 2006).
Liquefaction gained notice as an earthquake hazard in 1964 by causing major damage in
Alaska and Japan. Liquefaction from the March 27, 1964 Prince William Sound region Alaska
earthquake, of moment magnitude 3 (M) 9.2, caused widespread disruption to rail and road
networks (McCulloch and Bonilla, 1970) and may have contributed to a landslide that generated
a tsunami at the port of Valdez (Seed, 1968). Liquefaction from the June 16, 1964 Niigata
earthquake in northern Japan, M 7.5, led to the damage of the Showa Bridge and the tilting of
apartment buildings on reclaimed land adjacent to the Shinano River (see Figure 1.3). Earlier
earthquakes, such as the M 9.5 Valdivia earthquake of May 22, 1960 in Chile, and the San
Francisco, California earthquake of April 18, 1906 (M ~7.8), likely also resulted in liquefaction,
but the phenomena of liquefaction had not yet been recognized at those times. Liquefaction was
only identified in hindsight and therefore not extensively studied.
Damaging liquefaction has occurred during many subsequent earthquakes as well. In addition
to the damage to the Lower San Fernando Dam described above, liquefaction induced by the M
6.6 San Fernando earthquake in 1971 caused lateral deformations of up to 1.5 meters on gently
sloping ground over an area approximately 1,200 meters long and 270 meters wide (Youd, 1971;
Holzer and Bennett, 2007) and damage to a water treatment plant then under construction for the
California Water Project (Youd, 1973). The water treatment plant was to provide water for
Ventura County, West Los Angeles, Santa Monica, and the Palos Verdes Peninsula. Liquefaction
caused by the January 17, 1995 M 6.9 earthquake in Kobe, Japan (known as the Hyogo-ken
Nanbu earthquake), destroyed almost all of the deepwater berths at what was, at the time, the 6th
largest container port in the world. As a result, businesses and jobs went to ports in other
countries, and, in spite of repairs, the port only recovered to be the 17th largest port in the world
by 1997 (Chang, 2000) and the 28th largest by 2013.4 Liquefaction from the Christchurch, New
Zealand, February 22, 2011 M 6.1 earthquake produced lateral spreading, settlement, and sand
boils that devastated much of the city (see Box 1.2).

3
The moment magnitude scale is the magnitude scale most commonly used by seismologists and earthquake
engineers in current practice. Moment magnitudes provided throughout this report are those reported by the U.S.
Geological Survey (http://earthquake.usgs.gov/earthquakes/eqinthenews) and may differ from values reported in the
literature.
4
See http://www.worldshipping.org/about-the-industry/global-trade/top-50-world-container-ports.
20 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

(a) (b)

(c) (d)
FIGURE 1.3 Examples of liquefaction consequences. (a) Sand boils and ejected soil observed following
the M 6.5 Imperial Valley, California, earthquake in 1979. Note the pocketknife used for scale. SOURCE:
NOAA (b) The Showa Bridge collapse during the M 7.5 Niigata, Japan, earthquake in 1964. SOURCE:
NOAA; NGDC database. (c) The tilting of apartment buildings built on reclaimed land during the 1964
earthquake. SOURCE: NOAA; NDGC database. (d) Buoyant uplift of an underground gasoline tank
during the 2010 M 7.0 Darfield, New Zealand, earthquake. SOURCE: Courtesy of E. Kavazanjian.
INTRODUCTION 21

BOX 1.2
Recent Liquefaction in New Zealand

In addition to the 15,000 single family homes that had to be abandoned due primarily to liquefaction
associated with the February 22, 2011 M 6.1 earthquake (Rogers et al., 2014), many buildings in the
Central Business District also had to be demolished due to liquefaction-related damage. Figure 1 shows
the areas of liquefaction in the Christchurch area following that event. The event was part of a sequence
of several earthquakes beneath the Canterbury Plains (or greater Christchurch area) with an M of 5.0 or
a
greater, including three events greater than M 6.0. Up to 10 of those events are known to have induced
liquefaction (Quigley et al., 2013).
Several liquefaction-related issues were brought to the foreground by these events. For example,
the interrelationships between earthquake magnitude, distance, occurrence, and extent of liquefaction
can be demonstrated by comparison of the liquefaction induced by an M 7.0 event centered
approximately 40 km from Christchurch and the M 6.1 event that occurred directly beneath the city. The
larger event released approximately 22 times more energy than did the smaller event, but liquefaction
damage in Christchurch from the larger event was considerably less. Liquefaction induced by the smaller
earthquake was more extensive and damaging to Christchurch. The earthquake sequence also
demonstrated the potential for widespread damage to residential structures due to liquefaction (van
Ballegooy et al., 2014a) (see Figure 2); the potential for recurrence of liquefaction at a site; the potential
for liquefaction of silty soils; and the damage associated with widespread local occurrences of subsidence
and ejection of liquefied soil. Lessons regarding the impact of liquefaction on infrastructure and lifelines
(e.g., Cubrinovski et al., 2014; Kwasinski et al., 2014); on how liquefaction affects the performance of
foundations and buried pipelines (e.g., O’Rourke et al., 2014; van Ballegooy et al., 2014b); and on the
value of ground improvement in mitigating the consequences of liquefaction (e.g., Wotherspoon et al.,
2014) may also be drawn from this example.

FIGURE 1 Areas of liquefaction in Christchurch, shown in red, yellow, and pink, following the February
22, 2011, M 6.1 earthquake SOURCE: from NSF-sponsored GEER Report #GEER-027 dated 8 Nov.
2011
22 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

(a) (b)

FIGURE 2 Examples of the widespread damage to residential neighborhoods as a result of the 2011 M
6.1 earthquake in Christchurch. (a) Streets throughout the affected area were damaged due to
liquefaction (note displacement of pavement and sand ejected to the surface). 15,000 modern single
family homes, such as shown in (b), were damaged beyond repair and abandoned. SOURCE: photos
courtesy of Tara Hutchinson
____________
a
Values were derived from the USGS earthquake archive search of the ANSS Comprehensive Catalog
(http://earthquake.usgs.gov/earthquakes/search).

STATE OF PRACTICE FOR LIQUEFACTION ASSESSMENT

Across the United States, the state of practice for evaluating whether liquefaction can be
expected and what its consequences may be varies greatly. Factors influencing the state of
practice include the technical sophistication of the practitioners, their clients, and associated
regulatory agencies; the size of a project and the severity of consequences given the occurrence
of liquefaction; the type of project (e.g., dam, bridge, building, port and harbor facilities, or
nuclear power plant); and associated technical literature, guidance documents, standards, and
code provisions.
Higher levels of sophistication in liquefaction assessment, including site characterization
and analyses of consequences, might be expected in areas where liquefaction is commonly
recognized as a hazard, and among clients and regulators that work to mitigate the effects of
liquefaction more frequently. Higher levels of sophistication may also be typical of liquefaction
assessments for projects with major consequences given failure. Design guidance documents and
documents that describe best practices for liquefaction assessment have been prepared for some
specific classes of projects (e.g., for embankment dams [USBR, 2011], large dams [ICOLD,
2002], bridges [Dickenson, 2005; Kavazanjian et al., 2011, AASHTO, 2014], liquefied natural
gas facilities [Bachman et al., 2007], ports and harbor facilities [ASCE, 1998], and nuclear
power plants [USNRC, 2003]) as well as for general projects (CGS, 2008; SCDOT, 2010).
Nevertheless, the level of detail provided in such documents varies widely. In some cases, design
guidance may simply state that a liquefaction assessment should be conducted (OSBGE, 2014),
while in other cases detailed requirements for site investigation and a specific step-by-step
analysis procedure for liquefaction assessment may be provided (Kavazanjian et al., 2011).
Defining a standard of practice for liquefaction assessment is complicated by the rapid rate
at which recommended approaches are developed and modified. For instance, a technical paper
by Youd and colleagues (2001) describes techniques for assessment of the onset of liquefaction
(also referred to as triggering) developed by consensus at community workshops sponsored by
INTRODUCTION 23

the National Science Foundation and the National Center for Earthquake Engineering Research
at the State University of New York at Buffalo. By 2004, as a result of continued investigations
on the topic, alternative sets of changes to the consensus procedures were suggested by reputable
groups, with much emphasis placed by the technical community on two alternate approaches
(Boulanger and Idriss, 2004a; Cetin et al., 2004). Since then, liquefaction assessment in current
engineering practice has been inconsistent. Awareness of (and concern about) those
inconsistencies has been demonstrated in the results of a 2012 survey conducted by the Deep
Foundations Institute of engineers and contractors from the public, private, and academic sectors
(Siegel, 2013).

PRIOR REVIEWS OF LIQUEFACTION ASSESSMENT

Simplified approaches to estimate whether liquefaction is likely to be initiated (i.e., to


determine if the stiffness and strength of the soil may drop significantly due to liquefaction) were
proposed by Whitman (1971) and Seed and Idriss (1971). Most engineers have since relied on
variants of the Seed and Idriss (1971) basic approach (see Box 1.3). In 1985, the National
Research Council (NRC) held a workshop on liquefaction and issued a report that summarized
advances in liquefaction triggering assessments since the Alaska and Niigata earthquakes (NRC,
1985). In 1996 and 1998, the National Center for Earthquake Engineering Research (NCEER)
held two workshops on earthquake-induced soil liquefaction. The report issued following those
workshops (Youd et al., 2001) represents the most recent point at which a general consensus on
liquefaction triggering procedures was achieved. It is important to note that those workshops and
reports addressed liquefaction triggering only; they did not address issues associated with
liquefaction consequences.
Several groups have suggested alternatives to some of the NCEER workshop
recommendations (e.g., Zhang et al., 2002; Cetin et al., 2004; Bray and Sancio, 2006; Moss et
al., 2006, and Idriss and Boulanger, 2008). There can be substantial differences among these
revisions—relative to each other and relative to those in the NSF/NCEER workshop publications
(Youd and Idriss, 1997; Youd et al., 2001). The differences can be large, especially when
evaluating liquefaction triggering at depths greater than approximately 15 to 18 meters below
grade and in soils containing significant amounts of silt and clay particles (i.e., greater than 5%
by weight). Views differ on which revisions are most appropriate for liquefaction triggering
assessment. Further, as noted by Youd and colleagues (2001), there has never been consensus on
such issues as how best to assess liquefaction triggering on non-level ground (i.e., for slopes
greater than 6% or for embankments and dams) or on liquefaction consequence assessments.
Several alternatives to the simplified stress-based approach to assess liquefaction triggering
have been suggested. Strain-based approaches (Dobry et al., 1982; Dobry and Abdoun, 2011)
and energy-based approaches (Davis and Berrill, 1982; Berrill and Davis, 1985; Kayen and
Mitchell, 1997; Law et al., 1990; Green, 2001; Kokusho, 2013) have been proposed but are not
used widely in practice. Furthermore, liquefaction assessment has expanded beyond that of
triggering. It is no longer sufficient to assess whether liquefaction will occur at a site; it is
necessary to evaluate its potential effects on the strength, stability, and deformation of the ground
and the consequences on nearby structures and facilities. Various approaches to evaluate the
consequences of liquefaction have been developed over the past two decades (Ishihara and
Yoshimine, 1992; Martin et al., 2002; Youd et al., 2002; Ishihara and Cubrinovski, 2004;
24 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Kavazanjian et al., 2011). While development of analysis methods for liquefaction consequences
is a positive step, considerable uncertainty remains in the prediction of liquefaction triggering
and its consequences.

BOX 1.3
The “Simplified Method” for Assessing Earthquake-Induced Soil Liquefaction

A stress-based approach for evaluating whether liquefaction may be triggered at a site was originally
proposed by Whitman (1971) and Seed and Idriss (1971). While elements of this method, often referred
to as the “simplified method” (or the “Seed-Idriss simplified method”), have been subject to continuous
modification since 1971, the basic framework of this method is unchanged and it remains the most
commonly used approach to evaluate liquefaction triggering in practice today. In the simplified method,
the factor of safety (FS) against liquefaction triggering is defined as

𝐶𝑅𝑅
𝐹𝑆 =
𝐶𝑆𝑅

where CSR, the cyclic stress ratio, is a measure of the earthquake loading (the cyclic shear stress)
induced in the soil, and CRR, the cyclic resistance ratio, is a measure of the soil resistance (i.e., the cyclic
stress ratio expected to cause liquefaction). The “simplification” is to use the following equation, derived
from a simple application of Newton’s second law, to obtain a representative value of the CSR rather than
to conduct a detailed site response analysis:

𝑃𝐺𝐴 𝜎𝑣
𝐶𝑆𝑅 = 0.65 × × ′ × 𝑟𝑑
𝑔 𝜎𝑣𝑜

where PGA is the horizontal component of the peak ground acceleration at the site, g is the acceleration
of gravity, rd is a factor that accounts for the nonrigid response of the soil column, σv is the initial total
vertical stress in the ground, and σ’v is the initial vertical effective stress in the ground. To account for the
effects of the duration of earthquake shaking, which are related to earthquake magnitude, the CSR is
corrected by a magnitude scaling factor (MSF) such that

1
𝐶𝑆𝑅7.5 = 𝐶𝑆𝑅
𝑀𝑆𝐹

where CSR7.5 is an equivalent value of CSR for an M 7.5 earthquake.


In the original simplified method, the CRR was evaluated from the standard penetration test (SPT)
blow count of the soil, while subsequent variants of this method sometimes employ other indices of the in
situ soil resistance, including cone penetration test (or CPT) resistance and shear wave velocity (or Vs).
The correlation between the in situ soil resistance and CRR is typically presented using CRR7.5, the CRR
for an M 7.5 earthquake, for cases where there is no initial static shear stress on the horizontal plane or
2
effective overburden pressure of 96 kPa (I.0 ton/ft ). Correction factors are therefore applied to adjust the
CRR for overburden pressures different than 96 kPa (Kσ), and for an initial static shear stress on the
horizontal plane (Kα). The equation for FS can then be written as

(𝐶𝑅𝑅7.5 × 𝐾𝜎 × 𝐾𝛼 × 𝑀𝑆𝐹)
𝐹𝑆 = ⁄(0.65 × 𝑃𝐺𝐴 𝜎𝑣
′ × 𝑟𝑑 )
𝑔 𝜎𝑣𝑜

Much of the uncertainty associated with current practice for liquefaction analysis (and the significant
differences among the common methods used in practice to assess triggering) revolve around the value
of rd, the relationship between CRR7.5 and the soil resistance parameter (and in particular, SPT blow
count), and the value of the correction factors Kσ and Kα. Estimates for peak ground acceleration also
represent uncertainty. Chapter 4 of this report describes these issues in greater detail.
INTRODUCTION 25

THE COMMITTEE’S TASK

In 2010, an ad hoc committee of the Earthquake Engineering Research Institute (EERI)


called for a community-based evaluation of the state of the art in liquefaction assessment (Finn
et al., 2010). The EERI committee recommended that the assessment be conducted under the
auspices of the National Academies of Sciences, Engineering, and Medicine (the National
Academies) to help build consensus within the community and to lead to more confidence in
methods used to assess liquefaction potential and consequences. This report is the culmination
of the efforts in response to that recommendation from the EERI. Under the sponsorship of the
U.S. Bureau of Reclamation, Nuclear Regulatory Commission, Federal Highway
Administration, American Society of Civil Engineers, Port of Long Beach (California), Port of
Los Angeles (California), and the Los Angeles Department of Water and Power, the National
Academies convened a committee of 12 engineers and scientists to solicit input from the
technical community and to critically examine the state of practice in liquefaction hazard and
consequence assessment. The complete statement of task as agreed to by the National
Academies and the study sponsors is provided in Box 1.4.
The statement of task requires the committee to “assess and evaluate” various methods and
assessment approaches. The committee considered the difference between “assess” and
“evaluate” and determined the terms to be synonymous in this context. They are used
interchangeably throughout this report. Although the task requires the study committee to
evaluate objectively the various assessment methods (i.e., specific assessment procedures or
techniques), the task does not ask the committee to adjudicate disputes regarding any specific
methods or to recommend new best practices; it is not the intent of the committee to do so. The
committee has sought to identify and constrain, where possible, uncertainties in various
assessment methods and to identify what could be done in the near and long terms to advance
research and practice. Because many assessment approaches applied today are often
modifications of past approaches, and because much if not most of the data employed in more
contemporary approaches were collected prior to 1996, the committee both describes early
methods and databases to determine the uncertainties associated with them and also looks at
contemporary assessment approaches to determine how those uncertainties may have been
mitigated, exacerbated, or compounded. The report addresses evaluation of earthquake-induced
liquefaction only of naturally occurring soils and man-made fills composed of such soils. The
study scope does not extend to mine tailings, solid waste, or stabilized soils. Assessment of
remedial measures for increasing liquefaction resistance and mitigation of post-liquefaction
consequences is also outside the scope.
26 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

BOX 1.4
Statement of Task

An ad hoc committee of the National Academies of Sciences, Engineering, and Medicine will solicit
input from the technical community and critically examine the technical issues regarding liquefaction
hazard evaluation and consequence assessment. The study will assess and evaluate

 Sufficiency, quality, and uncertainties associated with laboratory and in situ field tests, case
histories, and physical model tests to develop and assess methods for determining liquefaction
triggering, and the resulting loss of soil strength and its consequences;
 Methods to conduct and analyze laboratory and physical model testing and to collect and analyze
field case history data to determine the triggering of liquefaction, and post-liquefaction soil
behavior (e.g., strength loss, dilation, and hardening); and
 Adequacy and accuracy of empirical and mechanistic methods to evaluate liquefaction triggering
and post-liquefaction deformations of earth structures and structures founded on or in the earth,
such as large embankment dams, levees, dikes, pipelines, highway embankments, bridges, pile-
supported decks, and other structural foundations. Effects at large depths and high static shear
stresses on liquefaction triggering and post-earthquake shear strength will be among those
addressed.

The study will focus on developments since the 1996 National Center for Earthquake Engineering
Research (NCEER) and the1998 National Science Foundation/NCEER workshops on liquefaction issues
and consider data including those related to soil properties, site characterization, ground motions, and
observations and measurements of soil response (e.g., post liquefaction deformations). Inherent
characteristics associated with the data (e.g., uneven distribution, scarcity, uncertainty) will be
investigated. The study will include a workshop on data gathering, vetting of field and laboratory data, and
new developments in the assessment of earthquake-induced soil liquefaction. The final report will assess
the state-of-the-art and practice for liquefaction analyses and will address future directions for research
and practice related to (i) collecting, reporting, and assessing the sufficiency and quality of field case
history observations as well as in situ field, laboratory, and model test data; (ii) addressing the spatial
variability and uncertainty of these data; and (iii) developing more accurate tools for assessing
liquefaction triggering and its consequences.

Committee Member Selection

Committee members were nominated by their peers and selected by the National
Academies from the resulting long list of nominees based on the strength of their professional
qualifications using long-standing procedures established by the National Academies to adhere
to the highest standards of impartiality and independence. The committee includes researchers
and practitioners with expertise in soil properties and behavior, in situ and laboratory testing,
ground motion propagation and attenuation, quantitative hazard and risk assessment, seismic
site response analysis, geological evidence of liquefaction, engineering geology, geotechnical
earthquake engineering, soil-structure interaction, and computational mechanics. Brief
biographies of the committee members are given in Appendix A.
INTRODUCTION 27

Information Gathering

Technical input from the engineering community was vital to the study process. Committee
conclusions and recommendations in this report are based on synthesis and assessment by the
committee of publicly available information and data. Additional information provided to the
committee by researchers during the conduct of this study was given with the understanding
that all information was subject to public disclosure. No primary research was conducted by the
committee. This report reflects information available to the committee as of early 2015.
Information made available later—for example, during the November 2015, 6th International
Conference on Earthquake Geotechnical Engineering in New Zealand—would have been
available too late for the study committee to deliberate its significance to this report and its
recommendations.
The committee organized and held a two-day public workshop with approximately 90
participants and held three additional public meetings and a webinar devoted to various topics.
Meetings were announced widely in advance through several communication mechanisms;
attendance at the workshop and public meetings was open to all interested parties; and remote
participation was possible for all of these sessions. Agendas for these meetings are provided in
Appendix B, as is a list of workshop participants. Solicitations for written input from any
member of the community on any relevant topic were made during all open session meetings.
The committee also requested written clarification on specific issues from individual members
of the technical community.

REPORT ORGANIZATION

Engineering practice related to earthquake-induced soil liquefaction is informed largely by


empirically based methodologies that are themselves informed by somewhat disparate
perspectives and uneven availability of data. Understanding the uncertainties associated with
various approaches and data is similarly uneven. Uncertainty, then, is an overriding theme in
this report because it is through an understanding of uncertainty that sources of inaccuracy are
identified. Consideration of uncertainties becomes even more important as earthquake
engineering practice more generally moves toward performance-based analysis and design.5
This report is organized as follows:

 Chapter 2 provides a description of the phenomena associated with earthquake-


induced soil liquefaction and the factors influencing them;
 Chapter 3 discusses the sufficiency of the case history data on liquefaction and
associated phenomena, including field case histories, and provides a critical
assessment of those data;

5
Performance-based design is a process in which the performance of a facility being designed is evaluated over the
entire range of possible loadings, rather than for one or more discrete ground motion return periods or events. The
approach is intended to allow rational decisions regarding appropriate design levels, including decisions related to
the added value of increasing the seismic resistance. Financial, construction, and operational issues are considered
over the life cycle of the facility. The approach usually involves probabilistic descriptions of earthquake ground
motions and the facility’s response to them.
28 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

 Chapter 4 describes and assesses the simplified stress-based approach to predict the
triggering (initiation) of liquefaction;
 Chapter 5 assesses alternative approaches to liquefaction triggering assessment such
as strain-based, energy-based, and computational mechanics-based approaches;
 Chapter 6 describes the assessment of the post-liquefaction shear strength of soils;
 Chapter 7 discusses empirical and semiempirical methods to evaluate liquefaction
consequences;
 Chapter 8 is a discussion of how computational mechanics can be used to predict
liquefaction triggering and consequences;
 Chapter 9 is a discussion of performance-based engineering methods for
probabilistic evaluation of liquefaction susceptibility, triggering, and consequences;
and
 Chapter 10 provides a synthesis of committee recommendations for advancing the
state of practice and state of the art for assessment of earthquake-induced soil
liquefaction.

Summaries of the key findings and conclusions derived from each chapter can be found in
a box at the beginning of the chapter, beginning with Chapter 2. A more detailed explanation of
each of the bulleted items in the key findings boxes is found within the text of the chapter.
These key findings and conclusions also inform the committee’s major recommendations,
found in Chapter 10.
2
A Primer on Earthquake-Induced Soil
Liquefaction

Key Findings and Conclusions

 Earthquake-induced soil liquefaction phenomena result from the interaction of soil particles and
porewater under the shear stress and shear strain reversals induced by earthquake shaking.
 Evaluation of liquefaction hazards requires assessment of the susceptibility of soils to liquefaction and
of the potential for liquefaction triggering and its consequences given anticipated ground motions.
 Not all soils are susceptible to liquefaction, but saturated, granular soils such as sands, silty sands,
low-plasticity silts and gravels have liquefied in past earthquakes.
 Liquefaction triggering depends on the level of ground shaking and the density and initial effective
stress of the soil; other factors such as soil type and age and the presence of pre-earthquake shear
stresses also influence the potential for triggering.
 Liquefaction can result in damage to buildings, bridges, lifelines, and other constructed facilities. In
many cases, the damage is a consequence of permanent deformation of the liquefied soil.
 In most cases, the in situ resistance of a soil to liquefaction currently cannot be predicted accurately
using laboratory testing. As a result, procedures developed to predict liquefaction triggering and
consequences rely on field observations documented in liquefaction case histories. These include
both empirical and numerical analyses, which need to be validated and calibrated using field data
before they can be considered reliable.

This chapter describes, in general terms for the less technical of this report’s readers, the
basic behavior of liquefiable soils under cyclic loading, including the effects of material type,
load amplitude and duration, soil density, initial effective stress, initial shear stress, and age of
the soil deposit. The susceptibility of various soil types (e.g., sand, gravel, and fine-grained soils)
to liquefaction triggering, the engineering measures of soil resistance to triggering, and the
primary consequences of triggering are also described. These descriptions are intended to
provide only a general overview; many topics in this chapter are expanded on in the chapters that
follow.

29
30 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

MECHANISM OF LIQUEFACTION

Understanding liquefaction triggering and its consequences requires an understanding of soil


and fluid mechanics. Liquefiable soils are frictional materials; their resistance to deformation is
influenced by how tightly the individual particles are pressed together. The term “effective
stress” (Box 2.1) is used to describe the stress associated with these interparticle contact forces.
Effective stress is defined as the external total stress on the soil less the internal porewater
pressure in the soil. It has long been recognized that soil behavior is governed by the effective
stress. Soils under high effective stress are generally stiff and strong, and soils under low
effective stress can be soft and weak.
Dry soil subjected to monotonic (i.e., unidirectional) shear loading may either decrease or
increase in volume depending on its initial density and initial effective confining stress and on
the levels of induced shear strain. Initially loose soil typically will tend to contract (i.e., decrease
in volume and become denser) as it is sheared. Under the same confining pressure, initially dense
soil will first contract but then dilate (i.e., increase in volume and become looser) as it is sheared.
On the other hand, when initially loose or dense dry soils are subject to repeatedly reversing
(cyclic) shear stress, they tend to decrease in volume, or contract, regardless of whether initially
loose or dense. When saturated soils are unable to contract due to water in the soil pores, the
water pressure increases. If it reaches the level of the initial effective stress, liquefaction can be
triggered. The extent to which a soil tends to contract or dilate during shearing dominates
liquefaction behavior.
Soils may be saturated (i.e., pore spaces are filled with water) or unsaturated (i.e., pore
spaces are filled with both water and air or just air). The degree of saturation (i.e., the fraction of
the pore space occupied by water) dominates liquefaction behavior. The degree of saturation
must be close to 100% for a soil to liquefy (Okamura and Soga, 2006; Yegian et al., 2007).

BOX 2.1
Effective Stress

Soils are an assemblage of individual particles (commonly with dimensions between 0.001 and 75
mm) that transmit applied stresses either through contact forces at particle contact points (see Figure 1)
or by a fluid (i.e., liquid or gas) pressure within the void space between the particles. It has been shown
that normal forces at particle contact points control volume change, stiffness, strength, compressibility,
and other important characteristics of the soil (Bishop, 1959; Skempton, 1960), but it is impractical to
account for all of those individual contact forces. Instead, an averaged applied normal stress is
determined by assuming that the components of the forces acting normal (i.e., perpendicular) to a given
plane have a well-defined average over the multiparticle scales for which the theory is to be applied. Then
an averaged interparticle normal stress (referred to as effective stress in soil mechanics and by
geotechnical engineers) is calculated by subtracting the fluid pressure within the void space from the
averaged applied normal stress.
Effective stress may be explained as follows. By force equilibrium, the total stress due to the weight
of soil particles, water, and any externally applied loads acting normal to any plane must be balanced by
the sum of the pore-fluid pressure and the interparticle forces acting on the plane divided by the area of
the plane. The effective stress principle states that the effective (interparticle) normal stress governs the
engineering behavior of soil. Annunciation of the effective stress principle by Terzaghi (1925) is generally
considered to be the birth of modern soil mechanics.
In a saturated soil, the pore-fluid pressure and the porewater pressure are the same. Therefore, the
effective stress may be written as ’ =  – u, where ’ is the effective normal stress,  is the total normal
stress, and u is the porewater pressure (see Figure 2). If the porewater pressure increases while the total
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 31

stress remains constant, the effective stress decreases. This reduction of effective stress is central to
triggering liquefaction.

FIGURE 1 Interparticle contact forces in a saturated soil. Soil stiffness and strength are governed by
these interparticle forces, but evaluating their magnitude is prohibitively expensive. Therefore, the
average normal stress on the plane of interest in the soil is used to represent the interparticle forces.
COURTESY: S. Kramer.

’ =  – u
FIGURE 2 Illustration of the effective stress principle. The effective stress on the horizontal plane is the
average intergranular normal stress on the plane. It is equal to normal stress on the horizontal plane
minus the porewater pressure. COURTESY: S. Kramer.

Behavior Under Monotonic Shear Loading

Monotonic shear loading refers to load conditions under which the shear stress (or principal
stress difference or deviator stress) or shear strain in the soil increases without change of
direction. Any increase in static load (e.g., as a result of foundation or embankment construction)
will induce monotonic shear loading in a soil. A decrease in soil strength (e.g., the strength loss
associated with liquefaction) can result in monotonic shearing deformation of a soil subjected to
an initial static shear stress (e.g., in sloping ground or beneath a foundation or an embankment).
When monotonically sheared, initially loose and dense dry sands deform in fundamentally
different manners with respect to their stress-strain and volume change (drained) behavior.
Figure 2.1 illustrates the stress-strain and volume change behavior (change in void ratio) of a dry
soil in loose and dense states confined under the same effective stress. The loose soil can be seen
to contract (decreasing void ratio) and the dense soil to dilate (increasing void ratio) with
increasing shear strain. Note that at large shear strain, neither loose nor dense soil continues to
change in volume (as shown in Figure 2.1b). At large strains both loose and dry soil reach the
same constant void ratio, termed the critical void ratio (ec). The critical void ratio is related
uniquely to the effective confining pressure via the critical void ratio (CVR) line (often called the
critical state line or steady- state line), as shown in Figure 2.2. The CVR line marks the boundary
between loose and dense states; that is, between where the soil demonstrates contractive and
32 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

dilative responses under monotonic shear loading. All soils tend to move toward the CVR at
large monotonic shear strains. Loose soil (i.e., soil with a void ratio greater than the CVR) is
contractive; its void ratio or effective stress must decrease to reach the CVR line. Dense soil (i.e.,
soil with a void ratio lower than the CVR) is dilative; its void ratio or effective stress must
increase to reach the CVR line.
If loads are applied to a saturated soil when there is not sufficient time for the water to move
in or out of voids (e.g., the load is applied rapidly or the soil has a low permeability), the loading
is said to be undrained. Since the water in a saturated soil is nearly incompressible, the tendency
for volume change will cause changes in porewater pressure and effective stress. In loose soils,
the tendency for contraction leads to increased porewater pressure and reduced effective stress.
In dense soils, the tendency for dilation leads to reduced porewater pressure and increased
effective stress.
Saturated soils with initial states that plot well above the CVR line are highly contractive
and, after reaching a peak shearing resistance at low strains, will generate high pore pressures
with concurrent large reductions in effective stress. Such soils can often shear to large strains
with low shearing resistance. Dense soils with initial states that plot well below the CVR line
will be dilative with decreasing porewater pressures and increasing effective stress; these soils
mobilize high shearing resistance at large strains when saturated due to decreasing pore pressure
and concurrent increase in effective stress. Soils of intermediate density (i.e., having density
between what is considered loose and dense) exhibit initially contractive behavior and mobilize a
peak shearing resistance at low strains followed by a reduction in shearing resistance. After
mobilizing the peak shearing resistance, however, these soils will dilate until a constant shearing
resistance is reached. These three conditions are illustrated in Figure 2.3 and are discussed in
more detail by Castro (1969).
A steady state of deformation, in which a soil shears with constant volume, constant
effective stress, constant shearing resistance, and constant strain rate, has been postulated to be a
unique function of void ratio and effective stress for a given soil (Castro and Poulos, 1977;
Poulos, 1981). Steady-state deformation can be represented graphically on a plot of void ratio
versus effective stress as a steady state line (SSL). To simplify the explanation, the SSL can be
considered nearly equivalent to the critical state CVR line. The degree to which a soil contracts
or densifies, therefore, depends on its state (i.e., its density and effective stress conditions)
relative to the SSL. The state parameter ( is a measure of how far the initial state of a soil plots
above (or below) the SSL and can be represented mathematically as

 = e – ess

where e is void ratio and ess is the void ratio under steady-state conditions. The value of
 reflects the influence of both mass density and effective stress on soil behavior, and it is an
index of a soil’s tendency to change volume. This tendency for volume change is, in turn, related
to the tendency for pore pressure and strength to change in a saturated soil when sheared.
Soils with positive state parameters plot above the SSL (see Figure 2.4b) and exhibit
levels of contractiveness (i.e., the tendency for pore-pressure increase and strength loss when
saturated) that increase with increasing state parameter value. Soils with negative
state parameters plot below the SSL and tend to be dilative (i.e., have a tendency for
pore-pressure decrease and strength gain when saturated) upon monotonic shearing. Under
cyclic loading, soils with both
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 33

positive and negative state parameters will tend to generate increasing pore pressure and can
liquefy under sufficiently strong loading.
The transition from contractive to dilative behavior observed in soils of intermediate density
occurs at the phase transformation point (Ishihara et al., 1975); that is, the point at which the soil
transitions from contractive to dilative behavior. The phase transition point represents a point of
minimum local shearing resistance. Undrained loading beyond the phase transformation point
produces dilation, a decrease in pore pressure, a higher effective confining pressure, and,
consequently, higher strength at large strains. The point at which a local minimum shearing
resistance is observed at moderate strain levels has been called the quasi-steady state of
deformation (Alarcon-Guzman et al., 1988). The shearing resistance at very large strain (i.e., at
the point where the shearing resistance no longer changes) is referred to as the ultimate steady
state (USS) strength (Yoshimine and Ishihara, 1998) to distinguish it from the quasi-steady state
strength.
As the void ratio of a soil decreases with increasing effective stress, the relationship
between the quasi-steady state line (QSSL) and ultimate steady state line (USSL) of a soil
changes with effective stress. Figure 2.5 illustrates the relationship between the quasi-steady
state and ultimate steady state at different initial effective stress levels (Yoshimine and Ishihara,
1998). The figure also illustrates the relationship of the isotropic consolidation line (ICL), the
line describing the relationship between void ratio and the effective confining pressure for one
particular soil prior to shearing the specimen. Because the ICL is typically flatter than the USSL,
the state parameter increases, and the soil becomes more contractive and potentially more
liquefiable, as the effective confining pressure increases.

FIGURE 2.1 (a) Stress-strain and (b) stress-void ratio curves for loose and dense sands at the same
effective confining pressure subject to drained loading. Note that the loose and dense soils eventually
converge on the same shear resistance and void ratio as shear strain increases. SOURCE: After Kramer,
1996.
34 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 2.2 Void ratio versus minor principal effective stress for deformation at constant volume (i.e., the
critical void ratio, or CVR, line) relationship separating dilative soil states from contractive soil states.
SOURCE: Kramer, 1996.

FIGURE 2.3 Schematic illustration of the undrained behavior of a sand in initially loose, dense, and
intermediate density states in undrained monotonic loading: (a) stress-strain behavior, and (b) effective
stress path behavior. Point marked with “x” indicates phase transformation point. Note that the ultimate
shear resistance of all three specimens falls on the “failure line” on the effective stress path plot, though
at different values of mean effective normal stress. SOURCE: After Kramer, 1996.

(a) (b)

FIGURE 2.4 (a) Behavior of initially loose and dense specimens under drained and undrained monotonic
loading conditions (with graphical illustration of state parameter for loose specimen) and (b) steady state
line with contours of equal state parameter as defined in Equation 2.1. SOURCE: Kramer and Stewart, in
preparation. COURTSEY: S. Kramer.
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 35

FIGURE 2.5 Stress-strain and stress path behavior for an element of soil consolidated to three different
initial states for testing. The paths show the transition of a soil from dilative to contractive as the effective
confining stress increases (after Yoshimine and Ishihara, 1998). Note that stress scales are not equal for
the three cases on the right-hand side; the ultimate steady strength would be highest for Case C and
lowest for Case A.

Behavior Under Cyclic Shear Stress Reversal

Earthquake loading may be characterized by repeated shear stresses of fluctuating intensity,


with the added characteristic that the direction of the applied shear stress reverses. Shear stress
reversal is an important characteristic of the earthquake loading, as both loose and dense soils
tend to contract at small induced shear strains and therefore generate positive excess porewater
pressures when subject to shear stress reversal (Martin et al., 1975). In steep slopes subjected to
weak shaking, the combination of high static shear stress and low cyclic shear stress can keep the
shear stress from reversing direction under undrained conditions, thereby preventing liquefaction
triggering. Cyclic shear stresses may also exceed the shear capacity of soils in sloping ground
and trigger a flow slide if the soil is loose of the critical state or contractive.
The stress-strain and pore-pressure generation behavior of sands subject to cyclic loading
have been investigated through laboratory tests. The stress-strain and stress path behaviors of a
clean sand1 subjected to undrained cyclic simple shear loading are described in Box 2.2. With
repeated stress reversal loading cycles, all liquefiable soils exhibit increases in excess pore
pressure (i.e., porewater pressure in excess of the initial steady-state pressure) and associated
effective vertical stress decreases. In liquefaction analyses, the excess pore pressure is often
expressed as a pore pressure ratio (ru)—defined as the ratio of the excess pore pressure to the
initial vertical effective stress. The pore-pressure ratio is an index of how close a soil is to
liquefaction. Prior to cyclic loading, ru = 0.0. As excess porewater pressure is generated, and thus

1
According to accepted engineering standards (ASTM, 2011), “clean” sands refer to those with less than 5% fines
content (i.e., grains smaller than 0.075 mm).
36 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

ru increases, the soil softens and the cyclic shear strain amplitude increases. The rate at which
softening and strain levels increase accelerates as liquefaction is approached. In some cases—
e.g., for sites with no initial static shear stress on the horizontal plane and for loose soil sites
where cyclic loading induces stress reversal—ru can approach and may actually reach 1.0 (at
which point the effective stress is zero). For sites without an initial static shear stress on the
horizontal plane, ru approaches and may actually be equal to 1.0 at liquefaction. For sites with an
initial static shear stress on the horizontal plane not subject to stress reversal, flow may occur (or
initiate) at values of ru less than 1.0. The softening and deformation produced by flow are driven
by the static shear stresses.

Effect of Initial Effective Stress

At a given density, steady-state and state parameter concepts predict that contractiveness
increases with increasing effective stress. Increased effective stress in the field is usually
accompanied by increased density (i.e., smaller void ratio), however, which tends to reduce
contractiveness. The cumulative effect of increased effective stress on contractiveness, therefore,
depends on whether the steady state line is steeper or flatter than the consolidation curve.
Typically, as shown in Figure 2.5, the steady state line is somewhat steeper than is the
consolidation curve. As a result, the cyclic stress ratio required to cause liquefaction decreases
with increasing initial effective confining pressure. Figure 2.6 presents the results of cyclic
simple shear laboratory tests performed without an initial static shear stress on the horizontal
plane on soils of the same density but different effective confining pressures (Vaid and
Sivathayalan, 1996). The CSR values in Figure 2.6 are all for the same number of cycles to
liquefaction. These curves illustrate the decrease in the cyclic shear stress ratio required to
trigger liquefaction in a given number of cycles with increasing vertical effective stress (i.e.,
increasing depth in the ground).
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 37

FIGURE 2.6 Curves presenting the cyclic stress ratio required to cause liquefaction (i.e., cyclic resistance
ratio, CRR) in simple shear tests after 10 cycles of loading illustrating the effect of increasing initial
vertical effective stress (and therefore depth) on liquefaction resistance (after Vaid and Sivathayalan,
1996).

BOX 2.2
Undrained Cyclic Laboratory Testing of Liquefiable Soil

The behavior of liquefiable soils such as clean sands subjected to cyclic shear loading with stress
reversals has been evaluated extensively in the laboratory for more than 40 years. Results of such tests
generally are not considered reliable for evaluating the liquefaction potential of soils in the field because
it is difficult to recover representative undisturbed samples. The tests have led, however, to better
understanding of the mechanics of liquefaction and the factors that affect liquefaction potential. The
figure below illustrates the behavior of a soil subjected to harmonic undrained loading in cyclic simple
shear without initial static shear stress on the horizontal plane. More detailed discussion of laboratory
testing methods can be found in Chapter 5. The following behaviors are illustrated:

 Pore pressure increases steadily (and vertical effective stress decreases) each cycle until the
effective stress path reaches the point at which the soil transitions from a contractive phase of
response to a dilative phase—often referred to as the phase transformation point (at cycle 21 in
Figures a.2 and a.4). Phase transformation points have been observed to fall on a unique line,
often referred to as the phase transformation line, for a given soil. The stiffness of the soil
decreases steadily and mildly up to the phase transformation point (Figure 1).
 With continued loading, the vertical effective stress becomes very low as the effective stress path
approaches the origin. It then becomes many times greater as the soil dilates and the effective
stress path moves up the failure envelope (Figure 2).
 The soil stiffness also increases and decreases within individual loading cycles (Figure 1) when the
phase transformation line is crossed. The stiffness becomes low when the effective stress is low
and then increases significantly as the soil dilates and the effective stress increases. This results in
banana-shaped stress-strain loops characterized by concave-up curves as the shear strain reaches
its peak within a loading cycle.
 The maximum increase or decrease in the shear strain during cyclic loading, the shear strain
amplitude, is small until the effective stress reaches a very low value (Figure 3). The shear strain
38 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

amplitude then increases quickly from cycle to cycle, even though the effective stress path does not
change. The rate at which the stiffness increases with increasing vertical effective stress at the end
of each banana-shaped stress-strain loop controls the strain amplitude in the later cycles of the
test.

These general patterns of behavior inform our understanding of the behavior of soil in the field as
liquefaction is approached. They provide the physical basis for constitutive stress-strain modeling of the
behavior of liquefiable soils discussed in Chapters 7 and 8.

(1) (2)
(1) (1)

(1) (1)

(3) (4)
(1)
(1)
(1)
(1)

FIGURE Illustration of behavior of typical loose liquefiable soil under cyclic loading. (1) Illustration of the
stress-strain behavior of an initially stiff soil (indicated by the steep curve). (2) Illustration of the effective
stress path during the test. Excess pore pressure increases with more cycles and the effective vertical
stress decreases. (3) Illustration of shear strain development prior to cyclic loading, ru = 0.0. After 22
cycles of loading, the effective stress path reaches the origin (i.e., ru  1.0), at which point the effective
stress is nearly zero and liquefaction has been triggered. (4) Effective stress reduction. Numbers
indicate cycle number (after Kramer and Stewart, in preparation). COURTESY: S. Kramer.

Liquefaction Susceptibility

Liquefaction susceptibility refers to the potential for liquefaction to be triggered in a soil.


Not all soils are susceptible to liquefaction, and the degree of susceptibility may depend on the
soil grain size and plasticity.2 The level of earthquake shaking and the in situ state of the soil

2
The plasticity of soil is the ability of a unit of soil to undergo permanent deformation under stress without
cracking. Plasticity is also an index of how changing the water content in a soil changes its behavior. Clayey soils,
generally considered more plastic than silty or sandy soils, can be deformed without cracking over a wide range of
water content. Tests to quantify plasticity have been standardized (ASTM, 2010). The level of plasticity is defined
using a plasticity index.
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 39

(e.g., density, soil fabric) at the time of shaking determine whether a susceptible soil will liquefy
in a particular earthquake. Factors that influence liquefaction susceptibility are reviewed briefly
and qualitatively herein. Determining susceptibility to liquefaction is discussed more
comprehensively in Chapter 4.
Liquefaction-susceptible soils generally are found within a narrow range of geologic
environments and soil age (Youd, 1991). Fluvial (i.e., river transported), colluvial (i.e., gravity
transported, such as at the base of a cliff), aeolian (i.e., wind transported), and other processes
that sort and deposit soils by grain size can result in the deposition of loose soils susceptible to
liquefaction. Liquefaction has occurred in alluvial-fan, beach, estuarine, and other deposits.
Many constructed earth fills (i.e., man-made soil deposits) are susceptible to liquefaction,
including, in particular, deposits created by hydraulic filling, which can replicate river and
stream processes.
Primary compositional factors that influence the liquefaction susceptibility of soils include
grain size and, for the finer-grained fraction, the plasticity. Grain size and plasticity are used to
identify soils not considered to be liquefiable (e.g., those that are fine grained and of medium to
high plasticity). Criteria for establishing whether or not a fine-grained soil is susceptible to
liquefaction are also addressed in Chapter 4.

FACTORS AFFECTING LIQUEFACTION POTENTIAL AND ITS CONSEQUENCES

A number of factors affect the potential for initiation of liquefaction. The next sections
include discussions on the effects of load amplitude, soil type, initial shear stress, shear strain
amplitude, age, and hydraulic conditions.

Effect of Load Amplitude, Duration, and Density

The excess pore pressure generated in a monotonic loading test on loose, saturated sand
increases with increasing shear stress, so the excess pore pressures under cyclic loading should
be expected to increase at a faster rate as the amplitude of the cyclic loading increases. An
increase in the duration of strong shaking should also be expected to increase excess pore-
pressure generation and the potential for triggering liquefaction. Figure 2.7 shows the
relationship between the number of uniform stress cycles required to induce liquefaction for a
soil at three different initial relative densities. The cyclic shear stress required to trigger
liquefaction (i.e., the CSR defined in Box 2.1), the number of cycles required to trigger
liquefaction, or both, for a given CSR can both be seen to increase with increasing density.
40 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 2.7 Cyclic strength curves used to characterize the resistance of a soil to liquefaction, showing
the influence of soil density (expressed as the relative density, Dr) on the variation of the number of
uniform stress cycles required to trigger liquefaction with the amplitude of applied loading (after De Alba,
P., H.B. Seed, and C.K. Chan. 1976. Sand liquefaction in large-scale simple shear tests. Journal of the
Soil Mechanics and Foundations Division 102(GT9):909–927. With permission from ASCE.).

Effect of Soil Type on Soil Behavior Under Cyclic Loads

The potential for a soil to liquefy during undrained cyclic loading is influenced by the extent
to which the soil skeleton tends to contract under such conditions. Loose, granular, and
nonplastic soils such as sands, gravels, and nonplastic silts can have high levels of
contractiveness. Therefore, porewater pressures within them can build, the soil softens and the
cyclic shear strain amplitude increases, and liquefaction will be triggered. Plastic fine-grained
soils may exhibit some softening under cyclic loads, but their tendency to contract is less, and
excess pore pressure may stabilize at values such that the effective stress does not approach zero
(and the soil does not liquefy).

Effect of Initial Shear Stress

No shear stress exists on horizontal planes at level-ground free-field sites (e.g., zero slope)
prior to earthquake shaking. In sloping ground, however, at level sites near slopes (e.g.,
riverbanks and embankment toes), and at level sites on which structural loads are imposed, there
will be an initial shear stress on the horizontal plane prior to earthquake shaking. That initial
horizontal shear stress can affect the rate of pore-pressure generation. Experimental studies (Seed
and Harder, 1990; Boulanger, 2003b) have shown that pore pressures are generated more quickly
in highly contractive (e.g., very loose) soils in the presence of an initial horizontal shear stress
than when no initial horizontal shear stress is present. In less contractive or dilative soils (e.g.,
medium dense to dense), however, the presence of an initial horizontal shear stress tends to
suppress pore-pressure generation. Laboratory studies have shown that the three-dimensional
stress state and the shear strain level can also affect pore-pressure development (Boulanger,
1990; Kammerer, 2002; Cetin and Bilge, 2015).
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 41

Effect of Shear Strain Amplitude

Different specimen preparation procedures produce soil specimens with different fabrics,
and differences in stress-based measures of cyclic shearing resistance observed in tests on
specimens reconstituted using different procedures (Ladd, 1974; Mulilus et al., 1977) led
investigators to explore the use of cyclic strain amplitude as a measure of loading, or demand, for
liquefaction. Data compiled by Dobry and Ladd (1980) suggest that the relationship between
pore pressure and cyclic shear strain is relatively insensitive to initial variations in the soil fabric.
That finding is consistent with the finding of Silver and Seed (1971): shear-induced volume
change is more closely related to cyclic shear strains than to cyclic shear stresses. Figure 2.8
shows the relationship between pore-pressure ratios and shear strain amplitude developed by
Dobry and Ladd (1980) from testing of reconstituted specimens of two soils, one of which was
prepared using three different specimen preparation methods (see Chapter 5).

FIGURE 2.8 Data for specimens prepared by different methods to Dr = 60% and subjected to 10 cycles of
uniform strain amplitude, suggesting a unique relationship between cyclic shear strain amplitude and
pore-pressure ratio (after Dobry, R., and R.S. Ladd. 1980. Discussion to “Soil liquefaction and cyclic
mobility evaluation for level ground during earthquakes,” by H.B. Seed and “Liquefaction potential:
Science versus practice,” by R.B. Peck. Journal of the Geotechnical Engineering Division 106(GT6):720–
724. With permission from ASCE).

Effect of Age

Empirical observations (e.g., Youd and Hoose, 1977; Youd and Perkins, 1978) indicate that
earthquake-induced liquefaction is more common in younger (e.g., Holocene age) soil deposits
than in older (e.g., Pleistocene age and older) soil deposits. Experimental investigation of age
effects is complicated by the difficulty simulating changes on geologic timescales, but
combinations of experimental and field data have shown that older soils are generally more
resistant to pore-pressure generation than younger soils are (Troncoso et al., 1988; Seed, 1979;
Arango et al., 2000; Robertson et al., 2000). Mechanisms of aging effects are not understood
completely, but they may involve particle and particle group reorientation, increases in in situ
lateral stresses, grain interlocking, chemical precipitation (cementation), and internal stress
arching (Mitchell and Solymar, 1984; Mitchell, 1986; 2008; Schmertmann, 1991) operating on
42 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

different timescales. It should be noted that aging effects can be “erased” when a soil’s structure
or fabric is disturbed or destroyed by events. Thus, the “age” of a soil deposit from a liquefaction
standpoint can be reset to zero by the triggering of liquefaction, even if large strains do not
develop (Andrus et al., 2009; Hayati and Andrus, 2009; Maurer et al., 2014a). The influence of
soil age on liquefaction potential is discussed in more detail in Chapters 3 and 4.

Hydraulic Considerations

Redistribution and dissipation of excess porewater pressure can strongly influence the
behavior of a soil as it approaches and undergoes liquefaction. Excess porewater pressure
generally develops in a spatially variable manner, producing hydraulic gradients that cause water
to flow from locations of high hydraulic head (i.e., locations of high excess porewater pressure)
toward locations of low hydraulic head (locations of lower excess porewater pressures).3 Water
from areas of high excess porewater pressure will often flow toward the ground surface. Sand
boils, the most common surficial evidence of liquefaction, may form when water and entrained
soil particles are ejected at the ground surface from a shallow liquefied soil layer (see Figure
2.9).
Excess porewater pressure migration and dissipation can be impeded if a low-permeability
soil layer lies above a liquefiable soil layer. In such cases, upward flowing porewater can build
up below the base of the low-permeability layer, and the density of the soil just below the low-
permeability layer can become very low, as illustrated in Figure 2.10. In extreme cases, a water
interlayer may form at the base of the low-permeability layer. Because porewater migration may
continue after the ground has stopped shaking, the most critical condition with respect to the
stability of a liquefied soil profile may exist after shaking ends.

FIGURE 2.9 Schematic illustration of sand boil and water interlayer formation (after Sims and Garvin,
1995). Water and entrained material may migrate from a liquefying layer through non-liquefying layers to
the surface.

3
The presence of an ambient artesian groundwater pressure could increase the rates and extents at which excess
porewater pressures develop, redistribute, and dissipate. A correlation between such pressure to liquefaction hazards
has been made between artesian pressures and liquefaction during the 2010 M 7.1 earthquake in New Zealand (e.g.,
Cox et al., 2012), although other researchers are less certain about the contributions (e.g., Tayler et al., 2012).
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 43

FIGURE 2.10 Illustration of void redistribution in which the volume of the liquefiable layer remains
constant but density changes within the layer. Note the loosening of the sand layer below the overlying
low-permeability clay layer, leading to a decrease in the strength of this layer (modified from NRC, 1985).

EARTHQUAKE LOADING

Liquefaction is triggered when the loading applied to a liquefaction-susceptible soil exceeds


the liquefaction resistance of the soil. Therefore, the evaluation of liquefaction triggering
requires establishing consistent measures of earthquake loading and liquefaction resistance.
Early investigations of liquefaction resistance were based on laboratory testing to develop cyclic
strength curves like those shown in Figure 2.7. Earthquake loading was, and still is, based on
representative values of the amplitude of the uniform cyclic shear stress and the number of
uniform loading cycles for the design earthquake. This approach requires development of
procedures to establish a representative earthquake-induced shear stress and the number of
uniform loading cycles as a function of earthquake magnitude, as discussed in the next section.

Ground Motion Intensity Measure

For most liquefaction analyses, the intensity of the design ground motion is characterized by
the peak cyclic shear stress and an earthquake magnitude. The peak cyclic shear stress, in turn,
depends on the design peak ground acceleration (PGA). As noted in Box 1.3, 65% of the peak
cyclic shear stress ratio is typically used as the value of the equivalent uniform cyclic load. This
reduced value of the peak cyclic shear stress was introduced to characterize the earthquake
loading to account for the fact that the peak cyclic shear stress occurs only once during the
earthquake and that all of the other load cycles during the earthquake are of lower amplitude.
With subsequently introduced magnitude scaling factors (MSFs) that account for durational
effects of the earthquake shaking, however, the use of the reduced (i.e., 65%) peak cyclic shear
stress ratio is no longer needed; nevertheless, it has been retained for historical purposes (i.e., use
of a different percentage reduction in peak cyclic shear stress ratio will result in either a shift in
the triggering correlation or a commensurate offset in the MSFs such that the cyclic stress ratio
adjusted to an M 7.5 will remain unchanged).
44 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Sources of the Design Intensity Measure

The earthquake intensity measure used in design (e.g., the peak ground acceleration) is
typically established based on a ground motion hazard analysis that considers the potential for
strong ground shaking at the site from recognized earthquake sources. Seismic hazard analyses
can be performed deterministically or probabilistically. Deterministic analyses consider a
particular scenario event, often taken as the largest assessed magnitude earthquake occurring at
the shortest source-to-site distance. Probabilistic analyses account for the distributions of
magnitude, source-to-site distance, and levels of ground shaking, and they compute the
probabilities of exceeding different ground shaking levels based on the probabilities of all
combinations of magnitude, distance, and ground shaking level. Project-specific hazard analyses
may be performed for certain projects, such as for critical infrastructure or in situations where
lives may be at risk. Alternatively, published results of seismic hazard analyses from a
recognized authoritative source (e.g., from the U.S. Geological Survey National Seismic Hazard
Mapping Program)4 may be used. The selection of an appropriate earthquake magnitude for use
with a probabilistic design acceleration is discussed in Chapter 4.

SOIL RESISTANCE TO LIQUEFACTION

The evaluation of liquefaction resistance has evolved in the last 40 years from laboratory-
based methodologies to a field-based framework, largely because obtaining high-quality samples
from the field for laboratory testing is difficult. The field-based framework characterizes
liquefaction resistance using a “representative” in situ test parameter from a test such as the
standard penetration test (SPT; described in Box 2.3), the cone penetration test (CPT; described
in Box 2.4), and shear wave velocity (Vs) testing (described in Box 2.5). Methods for evaluating
soil resistance to liquefaction are discussed in detail in Chapter 4.

4
See, for example, http://geohazards.usgs.gov/deaggint/2008/documentation.php.
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 45

BOX 2.3
Standard Penetration Test

For many years the standard penetration test (SPT) was the most common test for evaluating the in
situ consistency of soil, and it may still be very common in many areas. In the SPT, a standardized thick-
walled sampling tube (see Figure 1) is placed at the bottom of a borehole and driven into the ground by
means of a slide hammer with a standard mass and falling distance (see Figure 2). The number of
hammer blows required to push the tube into the bottom of the borehole for three six-inch intervals is
recorded. The SPT blow count, N, is the value of the sum of the number of hammer blows to penetrate
the second and third six-inch intervals. SPT procedures are codified in ASTM standard test methods
D1586-11 and D6066-11 (ASTM International, 2011a,b). The standard allows for a certain amount of
discretion in the methodology, however, which results in significant variability in SPT blow count
measurements. An advantage of the SPT is that it provides a soil sample that can be used for visual
classification, grain-size analysis, Atterberg limits determination, and other soil index properties.
Nevertheless, in addition to the uncertainties associated with blow count values, SPT measurements can
be limited if taken at widely spaced depth intervals (e.g., 0.75 to 1.5 meters). Features such as thin
potentially liquefiable soil layers may be missed.
Because of the lack of true standardization of the SPT, energy transmitted from the SPT hammer to
the tip of the SPT sampler is highly variable. Another major factor influencing blow counts is the effective
overburden pressure in the ground at the depth where the blow count is measured. As a result, stress-
corrected blow counts (N1), energy corrected blow counts (N60), and stress and energy-corrected blow
counts (N1)60 are often used in engineering practice (Kavazanjian et al., 2011). The various corrections to
measured SPT resistances are described in Chapter 4.

FIGURE 1 Cross section of the SPT Sampler. Note that the figure shows a sampler to be used with liners
to create a uniform inside diameter of 1.375 in. (34.9 mm). SOURCE: Coduto et al., 2011 (adapted from
ASTM D1586).
46 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 2 Illustration of one setup for a Standard Penetration Test (SPT) employing a hammer, rope, and
cathead per ASTM D 1586-11. Note that the figure shows a “donut” hammer with an average efficiency of
40%, requiring an energy correction factor of 0.67 to convert the measured value to N 60, the energy
corrected blow count. An additional correction for overburden pressure is required to convert N 60 to (N1)60,
the stress and energy corrected blow count often used in SPT correlations. SOURCE: Coduto et al., 2011
(adapted from Kovacs et al., 1981).

BOX 2.4
Cone Penetration Test

The cone penetration test (CPT) is an in situ test used to characterize the stratigraphy and
properties of soils. In the CPT, a standardized conical tipped probe is pushed into the ground at a rate on
the order of 20 millimeters per second. The force per unit area on the tip of the probe (referred to as the
tip resistance) and the force per unit area on the sleeve behind the tip (referred to as the sleeve
resistance) are measured as the probe is advanced into the ground. The CPT is codified in ASTM
standard D5778-12 (ASTM International, 2012). The CPT provides a nearly continuous (typically 5-10 cm
resolution) record of tip and sleeve resistance with depth that can be interpreted to identify thin layers and
small-scale fluctuations often missed during drilling, sampling, and SPT testing. CPT probes can be
outfitted with pressure transducers (piezocone or CPTu tests) to measure porewater pressures and with
geophones (seismic CPT) to measure shear wave velocities. A recent development is a cone-shaped
sampler that can be pushed by a CPT rig that can recover small soil samples for evaluation of soil index
properties (e.g., grain size). Cone penetrometers cannot be used in very stiff or dense soils, and they are
considered unreliable in soils containing gravel-sized and larger particles due to the difficulties in
penetrating them. The CPT is commonly used to investigate liquefaction case histories, but because the
CPT does not provide soil samples and the CPT-pushed cone-shaped sampler is not widely in use,
some drilling and sampling may still be required as part of a subsurface investigation. Figure 1 shows
schematic cross sections of typical electrical cone penetrometers, with two possible locations of the
porous filter. The most common CPTu probe is the one shown on the bottom half of the figure, with the
pore pressure transducer at the shoulder between the conical tip and the probe sleeve.
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 47

FIGURE 1 Cross-section of typical electrical cone penetrometers. SOURCE: Mayne, 2007.

BOX 2.5
Shear Wave Velocity

Measuring shear wave velocity (Vs) is another test used to characterize soils in situ. Vs refers to the
speed at which a shear wave (one type of wave generated by an earthquake) propagates through the
ground. The speed of wave propagation depends on the density of the soil, the directions of wave
propagation and particle motion, and the effective stresses in those two directions. Vs, by convention,
refers to the shear wave speed at very small amplitudes. Vs is related to the shear modulus of the soil at
small strain, Gmax, and the mass density of the soil, , by the equation:

𝑉𝑆 = √𝐺𝑚𝑎𝑥 /𝜌

where is equal to the total unit weight of the soil divided by the acceleration of gravity.
Vs measurements can be economical and non-invasive (i.e., there may not be a need to penetrate
the ground surface to make the measurement). The latter capability can be beneficial if soil profiles
contain inclusions (i.e., gravel or cobble inclusions) that can make testing difficult or even prohibit SPTs
and CPTs. There are many Vs measurement techniques, including downhole measurements (ASTM
International, 2014a), cross-hole measurements (ASTM International, 2014b), suspension logging (Nigbor
and Imai, 1994), and non-invasive methods (Stokoe and Santamarina, 2000). Because non-invasive Vs
tests do not provide soil samples, however, some drilling and sampling may still be required as part of a
subsurface investigation.

CONSEQUENCES OF LIQUEFACTION

The triggering of liquefaction can lead to various consequences to soil and site properties as
well as to physical damage, economic loss, and potential loss of life. Some factors that influence
the main consequences of liquefaction are described below. Descriptions of liquefaction
consequences and procedures for predicting those consequences are presented in Chapters 6 and
7.
48 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Alteration of Ground Motion

The response of buildings, bridges, pipelines, and other elements of infrastructure underlain
by liquefiable soils will be strongly influenced by how liquefaction affects the characteristics of
the ground surface motions. Ground motion frequency change often occurs suddenly when
liquefaction is triggered as a result of the rapid reduction in shear stiffness at high pore-pressure
ratios. The frequency content change can be so abrupt and obvious that it is easily noticed in
ground surface accelerograms.
While the onset of liquefaction typically reduces the intensity of the ground surface
acceleration (as compared to ground surface acceleration had the site not liquefied), the lower-
frequency components of the ground motion may increase subsequent to liquefaction triggering.
Even relatively low accelerations at low frequencies can produce very large, long-period
displacements following triggering—sometimes referred to as ground oscillations. Cases of
damaging ground oscillation have been reported in a number of earthquakes (Youd and Keefer,
1994; Youd, 2003; Holzer and Youd, 2007).

Lateral Spreading and Flow Sliding

Initial shear stresses under sloping ground conditions can drive permanent lateral
deformations of the ground subsequent to liquefaction. These lateral deformations are referred to
in practice as lateral spreading and flow sliding. When initial shear stresses are less than the
residual strength of the soil (i.e., the shearing resistance of the liquefied soil at large strain), but
the seismically induced stresses exceed the residual strength, lateral displacement builds up
incrementally during the earthquake and ceases when the cyclic loading stops and results in
lateral spreading. When initial static shear stresses are greater than the available strength of the
liquefied soil (e.g., in ground with a greater slope or lower soil density), lateral deformation will
not only build up during the earthquake but also continue to accumulate after the shaking stops.
This mechanism is referred to as a flow slide. While lateral spreading deformations may be
limited enough to be accommodated in engineering design, displacements caused by flow sliding
typically cannot be accommodated in design and must be addressed through mitigation
measures. Both lateral spreading and flow sliding are often accompanied by cracking of the
ground, separation between the ground and embedded structures, ejection of soil and water from
the ground cracks, and surficial settlement (e.g., formation of a graben) due to the lateral
movement. When the level of lateral spreading is small, these effects may be subtle and hard to
detect.
When earthquake shaking generates nonuniform excess pore pressures, hydraulic gradients
may develop that can, in turn, cause increased excess porewater pressures in other areas. This
would then result in a decrease in the effective stress and strength. This strength decrease can be
important for stability after the shaking has stopped. Figure 2.11 illustrates this phenomenon for
an embankment, but it can also occur beneath natural and man-made slopes. Slope failures
associated with this type of pore pressure redistribution have occurred from minutes to hours
after strong earthquake ground shaking (Seed et al., 1973; Ishihara, 1984). When porewater flow
is impeded by low-permeability soils, effective stresses can become extremely low and void
ratios can become extremely high. Such void redistribution effects can, in extreme cases, lead to
the development of water interlayers beneath the low-permeability zones.
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 49

FIGURE 2.11 Mechanism of delayed embankment failure due to redistribution of excess pore pressure
following earthquake shaking: (a) high pore pressure levels are generated below center of the
embankment where initial, static shear stresses are low; (b) following shaking, pore pressures increase
and effective stresses decrease in areas of high static shear stress (after Kramer and Stewart, in
preparation). COURTESY: S. Kramer.

Liquefaction-Induced Settlement

Volume loss due to the dissipation of excess pore pressure in the liquefied soil during and
after an earthquake generally results in ground surface settlement. Settlement can cause damage
to structures if the settlement is not uniform beneath the structure. Settlement can also leave a
gap beneath the pile-supported structures, cause distress to utilities buried in the soil, and alter
drainage grades at a site. The total volumetric strain (volume loss) following liquefaction is not
explained entirely by pore-pressure dissipation and the reconsolidation of the soil. This is
because cyclic loading of the soil beyond the liquefaction triggering changes the soil structure
such that the soil is more compressible than it was in its original state. Procedures for predicting
volumetric strain in saturated sands are discussed in Chapter 7. The settlement of structures
underlain by liquefiable soil is also affected by local interaction between the structure and the
soil, and it can differ significantly from the “free-field” settlement (Unutmaz and Cetin, 2012;
Bray and Dashti, 2014; Bray et al., 2014). Settlement of the ground surface can also be
associated with lateral soil movement and with the extrusion of soil (ejecta) from beneath the
ground surface.

Damage to Foundations

Liquefaction-induced damage to structures supported on shallow foundations is typically


caused by differential vertical and horizontal displacements. This is particularly true for
structures supported on isolated spread footings or lightly reinforced mats, as is often the case for
residential and light commercial structures (Cubrinovski et al., 2011; Bray et al., 2014).
Structures supported on stiffer mat foundations may also be subject to distortion, settlement, and
tilting when underlain by liquefiable soils. Embedded structures supported on mats or shallow
50 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

foundations may also be subject to damaging lateral loads and displacements associated with
lateral spreading and flow sliding.
Liquefaction-induced damage to deep foundations is associated typically with lateral
spreading and flow sliding. Lateral movement of the ground can induce lateral loads on pile caps
and bending moments and shear forces in the deep foundation elements. The magnitude of these
loads depends on the stiffness of the structure-foundation system and restraint provided by non-
liquefiable soils above and below the liquefied zone (see Figure 2.12). Settlement of liquefied
soil during and following an earthquake can lead to negative skin friction (i.e., downdrag loads)
on pile foundations as discussed in Chapter 7, although reports of damage to deep foundations
due to downdrag from the settlement of liquefied soil are rare.

FIGURE 2.12 Effects of lateral spreading of a liquefied soil with a non-liquefied crust on pile
displacement, d, and bending moment, M: (a) short embedment and thin crust, (b) deep embedment and
thin crust, (c) deep embedment and thick crust. The magnitude of the moment induced on the pile
depends on the thickness of the crust, the thickness of the liquefied layer, and the magnitude of the
lateral displacement; the worst case scenario is a thick crust laterally displacing above a thin liquefied
soil layer, as the thin liquefied layer localizes the bending in the pile within a narrow zone. COURTESY:
S. Kramer.

Damage to Retaining Structures

Bulging or tilting of retaining walls, sliding of the wall, and damage to structures on or
behind the wall may result if liquefaction-induced increases in earth pressure and decreases in
lateral resistance are not anticipated in design. In the 1995 Hyogo-ken Nanbu (Kobe, Japan)
earthquake, many gravity quay wall structures at the Port of Kobe settled, tilted, and slid laterally
due to a combination of increased lateral pressure, reduced lateral resistance, and bearing
capacity failure beneath the wall structures (see Figure 2.13). These deformations destroyed
gantry cranes that ran on rails along the tops of the walls and made the berths unserviceable,
resulting in significant economic loss to the port (see Chapter 1).
A PRIMER ON EARTHQUAKE-INDUCED SOIL LIQUEFACTION 51

(a) (b)
FIGURE 2.13 Damage to container cranes caused by liquefaction-induced lateral spreading of earth-
retaining structures in the 1995 Hyogo-ken Nanbu (Kobe, Japan) earthquake. Liquefaction-induced
damage rendered most of the berths in the port unserviceable, leading not only to physical damage but
also to loss of business that continues to impact the port to this day. SOURCE: NISEE, Pacific
5
Earthquake Engineering (PEER) Center, the University of California, Berkeley.

5
See https://nisee.berkeley.edu/elibrary/Image/K0124.
3
Case Histories

Key Findings and Conclusions

 The documentation and interpretation of liquefaction case histories have varied greatly in several
key respects, including
o distinction among case histories where liquefaction has or has not occurred, and where
evidence for liquefaction is ambiguous;
o identification of layers most susceptible to liquefaction and their representative properties;
o geologic controls on soil properties, particularly where layers vary laterally in shape and
content;
o estimation of the age and time since the last significant disturbance of liquefiable layers;
o estimation of seismic shaking to which layers were subject (i.e., seismic demand);
o identification of unsaturated zones below the groundwater table (e.g., resulting from
fluctuating groundwater levels); and
o consideration of the quality of the field data.
 More high-quality case histories are needed with parameter ranges outside those already well
represented in current databases, in particular for critical depths greater than approximately 15
meters and for earthquake magnitudes less than 5.9 and greater than 7.8, and including quality
case histories where no liquefaction was observed.
 Improved and truly standardized in situ test methods (e.g., standard penetration tests) and
protocols are needed to characterize relevant soil and profile properties, including thin layers,
anisotropy in soil fabric, and gravelly soils.
 Standardized protocols are needed for documenting and interpreting case histories of
liquefaction, lateral spreading and flow sliding, and related phenomena.
 Appropriate measures of the quality of case history data are needed in case history databases.
 More high-quality case histories are needed close to, and on either side of, the boundary lines on
plots that relate triggering to an in situ resistance parameter (e.g., normalized standardized
standard penetration test blow count).
 Field case histories need to be complemented with laboratory test data, particularly for
overburden pressures beyond the range constrained by field data, gravelly soils, and soils that
contain plastic fines as well as for stress conditions representative of sloping ground when
developing correlations that go beyond the bounds constrained by the field data.
 New instrument arrays are needed at sites where liquefaction is likely to occur in coming
decades. Such arrays should include surface and downhole instruments above and below
liquefiable layers and pore-pressure transducers within those layers.

52
CASE HISTORIES 53

Development of methods to assess liquefaction and its consequences relies on field case
histories. Geotechnical engineers have compiled databases that contain hundreds of these case
histories, the largest of which focus on whether liquefaction was triggered and the associated site
conditions. Other databases contain field examples of the consequences of liquefaction, such as
lateral spreading and flow slides. Those databases are important in developing and validating
methods for predicting liquefaction and its consequences (see Chapters 4-6).
A useful case history provides information about a soil’s in situ properties, its response to
seismic shaking, and the nature of that shaking. Ideally, the information would include in situ
properties prior to the earthquake and earthquake-induced ground motions measurements at the
site. Most often, however, soil profile characteristics are determined after they may have been
altered by the earthquake, and the seismic loading is estimated for the site rather than having
been recorded directly.
Most observations in the databases are of surface manifestations, not of the soil response at
depth (e.g., the extent of subsurface liquefaction and excess porewater pressure distribution). As
a result, use of case history data for development, calibration, or validation of liquefaction
analyses typically involves considerable judgment. Not surprisingly, varying interpretations of
the same field observations are not uncommon.
Despite these limitations, there is recognition within the profession of the value of well-
documented case histories. Post-earthquake surveys, including those under the auspices of the
National Science Foundation–sponsored Geotechnical Extreme Events Reconnaissance (GEER)
Association, provide important information before sites are altered by natural or anthropogenic
activities.
This chapter considers opportunities to strengthen the geologic and geotechnical basis for
describing and interpreting field case histories of liquefaction and related phenomena. The
opportunities described include enhancing site characterization with new technologies and using
new methods to improve case history protocols. The chapter focuses largely on case histories in
which an earthquake triggers, or fails to trigger, liquefaction. But the chapter also considers case
histories of two of the main consequences of liquefaction: lateral spreading, and the shear
strength of a soil that undergoes large deformation subsequent to liquefaction triggering.

SITE CHARACTERIZATION FOR CASE HISTORY ASSESSMENTS

One of the most important steps in documenting a liquefaction case history is to document a
site’s geologic and geotechnical conditions. In this site characterization, geologic descriptors
include tectonic setting, stratigraphy, landforms and land use, groundwater conditions, geologic
age, and prior earthquake history. Geotechnical observations include measurements related to
dynamic response characteristics of the soil and the liquefaction resistance of various deposits. In
liquefaction case histories, site characterization sets the stage for inferring whether and where
liquefaction has occurred within a profile and how its consequences evolved.

Geologic Characterization

Seismic shaking at a site depends on the fault-rupture location, mechanism, and dimensions;
on the rock structures through which the seismic waves pass; and on the geology directly beneath
the site. Whether and how extensively a saturated material liquefies can be influenced by
54 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

depositional environment, grain size distribution, lithology, weathering, age, and prior
earthquake history. Local stratigraphy and landforms affect how liquefaction at depth is
manifested at the ground surface.
Local geologic controls on liquefaction were recognized at least a half century ago, when
railroad damage from the March 27, 1964, Prince William Sound region Alaska earthquake (M
9.2) was related to “the underlying physiography and the materials of which the landforms are
composed” (McCulloch and Bonilla, 1970, p. D95). Studies of site geology have been important
in the interpretation of liquefaction case histories from several subsequent earthquakes in
California, including the November 24, 1987 earthquakes (M 6.5 and M 6.7) on the Superstition
Hills fault west of Westmorland (Holzer and Youd, 2007) and the October 17, 1989, Loma Prieta
earthquake (M 6.9) (Bennett and Tinsley, 1995; Boulanger et al., 1998; Mejia, 1998), and 1994
Northridge (Holzer et al., 1999).
Lateral variability in stratigraphic units can complicate the geologic and geotechnical
characterization of a case history site, and the lateral variability of an alluvial fan or meandering
riverbed deposits may differ significantly from those in man-made hydraulic fills. For example,
in one Loma Prieta earthquake liquefaction case history, the earthquake triggered liquefaction in
a channel fill but not in the sand just outside the channel fill (Bennett and Tinsley, 1995).
Similarly, the 2010 Darfield earthquake preferentially liquefied the fills of abandoned river
channels in Christchurch, New Zealand (Wotherspoon et al., 2012).
Protocols for documenting geologic site conditions could include attention to what
geologists call sedimentary architecture—the geometries of the various types of deposits laid
down by water and wind. Geologic clues to these geometries can be combined with surface and
subsurface investigations to aid estimating the extent of liquefiable layers through interpolating
between and extrapolating beyond the places where geotechnical observations are available.

Geotechnical Characterization

Most liquefaction case histories include data from field tests for measuring soil properties. In
situ testing avoids disturbance of the soil from sampling, particularly for nonplastic soils, and it
obviates the need to reestablish the in situ state and fabric in the laboratory. Three in situ
methods used most widely to evaluate soil properties for case histories are the standard
penetration test (SPT; see Box 2.4), the cone penetration test (CPT; see Box 2.5), and shear wave
velocity (Vs) testing (see Box 2.6). Table 3.1, updated from Youd and colleagues (2001),
summarizes the advantages and disadvantages of these tests for assessing liquefaction triggering
potential. The main advance in liquefaction potential assessment since Table 3.1 was originally
published is the increased size of the Vs liquefaction triggering database. The liquefaction
triggering database now contains more case histories with documentation of Vs than of either
SPT or CPT resistance (Kayen et al., 2013). Nonetheless, the number of documented case
histories using SPT and CPT has also increased since 2001, and more importantly, some of the
databases include quality ratings and parameter uncertainties assigned to their case histories. The
number of high-quality case histories continues to increase with recent additions from
earthquakes offshore Bio-Bio, Chile (February 27, 2010; M 8.8), near Honshu, Japan (March 11,
2011; M 9.1), and the series of earthquakes in or near Christchurch, New Zealand in 2010 and
2011.
Other in situ test methods used to characterize soil behavior and stratigraphy for liquefaction
case histories include the piezocone penetration test (CPTu). The CPTu can be particularly useful
CASE HISTORIES 55

for identifying dilatant zones, providing information regarding groundwater elevations, and
identifying very thin silt and clay seams missed by other in situ test methods (Mayne, 2007).
More esoteric techniques include an electrical probe to assess the anisotropy in soil fabric
(Arulmoli et al., 1985; Arulmoli and Arulanandan, 1994; Arulanandan, 2008) and the vision
cone penetrometer test (VisCPT) to identify dilatant zones, to provide information regarding
groundwater elevations, and to identify very thin silt and clay seams missed by other in situ test
methods (Mayne, 2007). The VisCPT may also provide information on fines content (Raschke
and Hryciw, 1997). These methods need to be more fully developed, however, and their efficacy
demonstrated to add value for documenting and interpreting case histories, before they are
embraced by the engineering community.
The presence of coarse gravel-sized and larger particles in soil requires use of methods other
than the SPT and the CPT. Large particles physically interfere with the advancement of the SPT
sampler and the CPT cone into the ground, resulting in measured resistance values that are too
high. The main in situ methods used for gravelly soils are the Becker penetration test (BPT)
(Harder and Seed, 1986), the Chinese dynamic penetration test (DPT) (Cao et al., 2013), and
various large penetration tests (LPTs) (Daniel et al., 2003). BPT and LPT efforts have mainly
focused on developing correlations relating their penetration resistance to that of the SPT. These
equivalent SPT soil resistance correlations are then used to assess liquefaction triggering
potential. Recent advances have been made in the development of improved SPT-BPT
correlations by instrumenting the BPT to directly measure the energy transferred from the
hammer to the driving shoe, avoiding the need to estimate the energy loss due to the frictional
resistance between the drill string and the surrounding soil (Ghafghazi et al., 2014; Dejong et al.
2014). Efforts with the DPT have focused on direct correlation between DPT penetration
resistance and liquefaction resistance. The DPT case history database is relatively small, though,
and the quality of the cases has not been formally assessed.
Given that other methods are less mature, Vs testing may provide the most reliable data for
evaluating liquefaction triggering resistance in coarse-grained soils not amenable to use of the
SPT or the CPT. Development of the database of case histories of Vs measurements in coarse-
grained soils in which liquefaction triggering is suspected should continue. In addition, however,
the instrumented BPT and the DPT both show significant promise, and their further development
would benefit the technical community.
There are far fewer case histories that document in situ soil properties at sites where lateral
spreading and flow sliding have occurred than are found in liquefaction triggering databases. The
SPT has been the primary method of characterizing the former phenomena, where these case
studies have been used to develop correlations to estimate the post-triggering strength of the
liquefied soil. Of the few case histories of lateral spreading or flow sliding characterized by the
CPT, many of the data points are from older and less reliable CPT probes. In the residual shear
strength of liquefied soil case history database compiled by Robertson (2010), only 6 of the 34
case history sites were characterized using the modern electronic CPT. Data collection for
residual shear strength of liquefied soil databases need to include testing with modern CPT
probes and other new and improved in situ test methods to enrich the available databases.
Triggering may also be identified with additional investigation, including a search for dikes and
sills underground, as illustrated in Box 3.1.
56 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

TABLE 3.1 Advantages and Disadvantages of Various Field Tests for Assessment of Liquefaction
Triggering
Test
Attribute SPT CPT Vs
a
Cumulative use in liquefaction case histories Abundant Abundant Abundant

Stress-strain conditions during test Partially drained, Drained, large Small strain
large strain strain

Quality control and repeatability Poor to good Very good Good

Detection of vertical stratification of soil Good for closely Very good Depends on
deposits spaced tests testing
procedure, but
generally fair to
b
good

Soil types in which test is recommended (i.e., All but gravelly All but gravelly All, with due
is expected to provide a reliable index of soils soils consideration of
liquefaction resistance) particle size
c
effects

Soil sample retrieved Yes No for standard No


d
configuration

Test measures index or engineering property Index Index Engineering


property
a
Kayen et al. (2013).
b
Stokoe and Santamarina (2000).
c
Committee comment.
d
Robertson and Cabal (2014).
NOTE: Committee updates are in italics.
SOURCE: Modified from Youd, T.L., I.M. Idriss, R.D. Andrus, I. Arango, G. Castro, J.T. Christian, R. Dobry, W.D.L.
Finn, L.F. Harder, M.E. Hynes, K. Ishihara, J.P. Koester, S.C.C. Liao, W.F. Marcuson, III., G.R. Martin, J.K. Mitchell,
Y. Moriwaki, M.S. Power, P.K. Robertson, R.B. Seed, and K.H. Stokoe, II. 2001. Liquefaction resistance of soils:
Summary report from the 1996 NCEER and 1998 NCEER/NSF Workshops on Evaluation of Liquefaction Resistance
of Soils. Journal of Geotechnical and Geoenvironmental Engineering 127:817–833. With permission from ASCE.
CASE HISTORIES 57

BOX 3.1

Subsurface Evidence for Liquefaction

Sand boils and their underground feeders are valuable indicators of past earthquakes (Obermeier,
1996). Liquefaction is known to recur at the same location in successive earthquakes (Youd, 1991; Tohno
and Shamoto, 1986; Yasuda and Tohno, 1988; Sims and Garvin, 1995; Quigley et al., 2013). Liquefaction
features have provided the main basis for estimates of earthquake probabilities in areas such as
Charleston, South Carolina (Talwani and Schaeffer, 2001), and the New Madrid seismic zone of
Arkansas and Missouri (Tuttle and Hartleb, 2012).
Outcrops, trenches, and borings can reveal traces of sand boils, sand intrusions, and other evidence
for fluid escape (see Figure 1). Water ejected from liquefied soils, with or without solid material from those
soils, can create or follow underground pathways that are marked by intrusions whether or not the water
produces sand boils at the ground surface (Lowe, 1975; Hurst, et al., 2011). The intrusions are termed
“dikes” if steep and “sills” if close to horizontal. Methods to identify them include making sediment peels
that bring out bedding and its disruption (see Figure 2) (Nakata and Shimizaki, 1997, 2000; Takada and
Atwater, 2004).

FIGURE 1 Sand intrusions exposed in a drainage ditch near New Madrid, Missouri. The dike branches to the left and
right into sills that run beneath a cap of weathered clay. The intrusions are interpreted to represent no fewer than two
earthquakes. The scale in the lower half of the photo is 8 cm. SOURCE: M. Tuttle
(http://mptuttle.com/newmadrid4.html).
58 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 2 Rectangular cores 0.5 m wide and 8 m deep revealed dikes and sills in deposits less than 2,000 years old
beneath banks of the lower Columbia River. The intrusions were found beneath banks that abound in exposed dikes,
but they were also found beneath banks where this surficial evidence for liquefaction is scarce or absent. Fluid
escape from sand less than 2,000 years old created sills beneath mud lenses and dikes that cut through those lenses
but which dissipate in bedded sand above. SOURCE: Modified from Takada, K., and B.F. Atwater. 2004. Evidence
for liquefaction identified in peeled slices of Holocene deposits along the lower Columbia River, Washington. Bulletin
of the Seismological Society of America 94:550–575 © Seismological Society of America.

LIQUEFACTION TRIGGERING CASE HISTORIES

The largest databases for liquefaction triggering are those assembled for use in developing
the simplified, stress-based liquefaction evaluation procedures discussed in Chapter 4. Central to
these procedures is use of the cyclic resistance ratio (CRR). In a simplified procedure, CRR is
correlated to one of the in situ test parameters described above (e.g., SPT blow count, CPT tip
resistance, or Vs). The evolution of liquefaction triggering case histories compiled to develop
CRR curves is summarized below. The databases used to develop the most common triggering
curves, as well as the challenges in interpreting and the opportunities to improve liquefaction
triggering case histories, are also discussed.

Overview of Triggering Databases

Published collections of triggering case histories date back to at least 1971. Whitman (1971)
assembled 13 different cases from 8 earthquakes in Chile, Japan, Mexico, the Philippines, and
the United States. Entries for each case included the depositional environment of the soil inferred
to have liquefied; the estimated peak ground acceleration (PGA) at the ground surface; the depth
of the groundwater table; the depth of the soil inferred to have liquefied (i.e., the “critical layer”);
a representative SPT blow count (or N-value) for the critical layer; and the estimated duration of
strong earthquake shaking. Seed and Idriss (1971) compiled 35 cases from 12 earthquakes in
Chile, Japan, and the United States. Entries in that database included earthquake magnitude,
CASE HISTORIES 59

distance from the earthquake source, and soil type in addition to water-table depth, critical-layer
depth, representative SPT N-value for the critical layer, and estimated values of PGA and strong
shaking duration. The Seed and Idriss (1971) database included 23 cases where liquefaction is
known to have occurred and 12 cases in which no liquefaction is inferred.
Many other liquefaction triggering databases have been assembled since 1971, chiefly to
support the development and updating of CRR curves. The databases vary in their levels of
documentation. Examples listed chronologically and grouped by in situ test method include:

Standard penetration test (SPT):Yegian and Vitelli (1981), Tokimatsu and Yoshimi (1983),
Seed et al. (1984), Liao and Whitman (1986), Cetin et al. (2000), Idriss and
Boulanger (2008), Boulanger and Idriss (2014);
Cone penetration test (CPT): Stark and Olson (1995), Suzuki et al., (1995), Olson and Stark
(1998), Moss (2003), Idriss and Boulanger (2008), Boulanger and Idriss (2014);
Shear wave velocity (Vs) testing: Andrus and Stokoe (1997), Andrus et al. (2003), Kayen et
al. (2013).

Other liquefaction triggering databases have been compiled in support of the development of
CRR curves for other types of in situ tests, including the flat plate dilatometer test (Monaco et
al., 2005) and the Chinese dynamic penetration test (Cao et al., 2013).
Table 3.2 summarizes the databases used to establish five liquefaction triggering curves
used commonly in engineering practice, grouped by the in situ test method used to characterize
soil resistance. The table reports the number of data points for cases in each database where (1)
liquefaction is known to have occurred; (2) liquefaction is not believed to have occurred; and (3)
borderline cases (“yes/no” cases). Other attributes in this table include the ranges of the depth to
the center of the layer, with the lowest factor of safety against triggering (i.e., the critical layer)
and the effective overburden pressure associated with the critical layer; the fines content (percent
by weight passing the #200 sieve); and the normalized value of the penetration resistance or
small strain Vs (N1,60cs, qc1Ncs, Vs1) for each database.
Figure 3.1 illustrates the range of the normalized value of the penetration resistances and
small strain shear wave velocities for three of these databases (one for the SPT, one for the CPT,
and one for Vs). Figure 3.1 indicates that the case history database is deficient in case histories
for depths greater than 12-15 meters depending on the in situ test method; for normalized and
standardized clean sand blow count, N1,60cs, in excess of 25; for normalized CPT tip resistance,
qc1Ncs, greater than 125 atm (12.7 MPa); and for normalized Vs (Vs1) in excess of 225 m/s.
Histograms of the ranges of the various parameters tabulated in Table 3.2 for the recently
assembled SPT, CPT, and Vs databases are shown in Appendix C. Those histograms also suggest
that databases may be deficient with respect to magnitudes less than 5.9 and greater than 7.8-8.2,
again depending on which in situ test is under consideration.
A significant issue when evaluating these databases is the criteria used to determine the data
points and attributes summarized in Table 3.2. Subjective assessments of which data points to
include for the “yes,” “no,” and “yes/no” cases and selection of their associated critical layers
have been applied during database development. Issues associated with these assessments are
described in the next section. Given the importance of data point and attribute selection in the
development of CRR triggering curves, standardized protocols need to be applied in the
development of future databases. There is some movement in this direction within the profession
(e.g., Stewart et al., 2015), but no liquefaction case history database has been compiled to date
60 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

using standardized procedures developed by consensus within the profession. Where possible,
such procedures need to emphasize use of multiple in situ test techniques to characterize the soil
layer expected to have liquefied, including primary wave (p-wave) velocity measurements to
confirm saturation of the layer (e.g., for soil profiles where the groundwater table fluctuates
regularly or, possibly, where decaying organics may release gases into the soil layer that result in
unsaturated conditions). This will facilitate comparison and correlation among the different in
situ test techniques.

TABLE 3.2 Summary of Recently Compiled Liquefaction Triggering Case History Databases for Level-
Ground Conditions Showing Ranges in Values of the Parameters
SPT CPT Vs
a b c d
Parameter Cea04 BI14 Mea06 BI14 Kea13
“yes” cases 287
109 133 139 180
“no” cases 124
88 118 44 71
“yes/no” cases 4
3 3 0 2

Critical depth (m) 1.1-20.5 1.8-14.3 1.4-14.0 1.4-11.8 1.1-18.5

Effective overburden
stress’vo (kPa) 8.1-198.7 20.3-170.9 14.1-145.0 19.0-147.0 11.0-176.1

Fines content (% by
weight) 0-92 0-92 - 0-85 -

N1,60cs (blows/30 cm),


qc1Ncs (atm), or Vs1 e
2.2-66.1 4.6-63.7 11.2-252.0 16.1-311.9 81.7-362.9
(m/s)

Cyclic stress ratio


f
CSRM7.5 0.05-0.66 0.04-0.69 0.08-0.55 0.06-0.65 0.02-0.73

Earthquake
5.9-8.0 5.9-8.3 5.9-8.0 5.9-9.0 5.9-9.0
magnitude M
a
Cea04: Cetin et al. (2004).
b
IB14: Boulanger and Idriss (2014).
c
Mea06: Moss et al. (2006).
d
Kea13: Kayen et al. (2013).
e
N1,60 values listed for Cea04, as opposed to N1,60cs.
f
CSR values listed for Mea06 and Kea13, as opposed to CSRM7.5.
CASE HISTORIES 61

(a) (b)

(c)
FIGURE 3.1 Plots of normalized in situ test values for case histories as a function of depth for the most
recently compiled databases: (a) SPT (Boulanger and Idriss, 2014); (b) CPT (Boulanger and Idriss,
2014); and (c) Vs (Kayen et al., 2013). COURTESY: R. Green
62 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Liquefaction of Fine-Grained Soils

Soils with greater than 50% of their constituents passing the #200 sieve (i.e., fine-grained
soils with grain sizes smaller than 0.075 millimeters) were once considered non-liquefiable, but
several investigators have reported liquefaction of fine-grained soils in earthquakes. Wang
(1979), as cited by Bray and colleagues (2004a,b) and Boulanger and Idriss (2004b), reported on
liquefaction of “silty soils” in seven earthquakes in China. Based on data provided in Wang
(1979), Seed and Idriss (1982) proposed criteria for assessing the susceptibility of fine-grained
soils to liquefaction, commonly referred to as the “Chinese criteria.” These criteria are a function
of percent clay (less than 15% by weight of particles < 0.005 mm), liquid limit (LL < 35%), and
with an in situ water content greater than 0.9 multiplied by the LL. Subsequent studies, however,
prompted by field observations of the cyclic response of fine-grained soils following more recent
earthquakes, raised questions about the validity of the Chinese criteria (e.g., Seed et al., 2003;
Bray et al., 2004a, b; Bray and Sancio, 2006). Those studies indicated that the percentage of clay
was less important than was the plasticity index (PI) of the soil and that the ratio wc/LL was
more important than was the LL alone in assessing the liquefaction susceptibility of fine-grained
soils. Furthermore, the characteristics of the cyclic behavior of the soil and whether the soil’s
strength and compressibility were more aligned with clay than sand were identified as being
important in evaluating the potential for ground failure of fine-grained soil deposits during
earthquakes (Boulanger and Idriss, 2006). In the absence of more detailed laboratory or in situ
data, PI alone of the fine-grained soil was judged by Boulanger and Idriss (2006, 2008) to be a
suitable proxy for determining whether the soil behaved as “clay-like” or “sand-like.”
Although the recommended PI threshold among the various studies (i.e., Boulanger and
Idriss, 2004b, 2006, versus Bray et al., 2004a,b, and Bray and Sancio, 2006) at first appearance
seem inconsistent, in reality the purpose of the recommendations made in these studies differs.
Specifically, Bray and Sancio (2006) provide the results from a detailed laboratory study focused
on determining the characteristics of fine-grained soils that are and are not susceptible to
liquefaction. In contrast, Boulanger and Idriss (2008) focus on the more pragmatic issue of
whether the simplified liquefaction evaluation procedure is suitable to evaluate the liquefaction
potential of a fine-grained soil or whether laboratory testing needs to be performed. Seed and
colleagues (2003) attempt to address both the fundamental issue of the susceptibility of fine-
grained soils to liquefaction and the pragmatic issue of how to evaluate soil liquefaction
potential.
All of the studies mentioned above focused on soils with relatively high fines contents (i.e.,
fines contents above that required to fill the voids of the coarse-grain fraction of the soil,
resulting in the coarse grains “floating” in the fine-grain matrix). As a result, there is uncertainty
about the applicability of the recommendations from these studies for assessing liquefaction
susceptibility of soils having lower fines contents. Furthermore, interpretation of published field
data is complicated by the general failure of the investigators to report Atterberg limits for fine-
grained soils that were subject to strong shaking but did not liquefy. Additional field case history
data on the performance of fine-grained soils subject to strong shaking, including Atterberg
limits on soils that did not liquefy, supplemented by sampling and laboratory testing of the fine-
grained soils reported to have liquefied and not liquefied, are needed to develop comprehensive
criteria regarding the liquefaction susceptibility of fine-grained soils and the suitability of various
procedures for evaluating liquefaction potential. These data needs include soils with liquid limits
CASE HISTORIES 63

greater than 35% and PIs greater than 15%, which, to date, represent the limits of soils reported
to have liquefied in the field.

Interpretative Issues in Liquefaction Triggering Case Histories

To determine whether or not soils at a site have liquefied, and which deposits are likely to
have liquefied if liquefaction is inferred to have occurred, representative geotechnical properties
must be established for the soils. More often than not, these properties are determined using data
collected after an earthquake, despite the fact that properties may have changed as a result of the
earthquake. Earthquake ground motions at case history sites are usually estimated based on
nearby recordings, event-specific contours of strong ground motions, or generalized ground
motion prediction equations, as cases are rare where motions are recorded at a site known (or
suspected) to have liquefied.

Determining Whether Liquefaction Was Triggered

With a few exceptions, whether a case history site is classified as having liquefied or not is
based on the presence or absence of post-earthquake surface manifestations of liquefaction (e.g.,
sand boils and ejecta, cracking, or settlement). When surface manifestations of liquefaction are
observed, it is assumed that at least one deposit in the soil profile has liquefied. Changes to the
surface (or lack thereof) are not necessarily accurate indicators of liquefaction having occurred
(or not), however. For example, there may be settlements due to dissipation of excess porewater
pressures in sandy soils that occur when the factor of safety against liquefaction is upwards of
two, particularly in looser deposits (Ishihara and Yoshimine, 1992). In such a circumstance, it is
possible that case histories could be classified erroneously as having liquefied. Similarly,
liquefaction may not manifest itself at the ground surface due to the depth or density of the
liquefied layer or the integrity and thickness of the overlying non-liquefied crust. In most
databases, the “yes/no” cases are those with marginal or minor evidence of liquefaction.
Discrepancies in interpretation do occur even in identifying the occurrence of liquefaction: one
such example is Wood and colleagues (2011) versus Smyrou and colleagues (2011). And, there
is little to no consistency among databases in how to define those cases in which evidence of
liquefaction is ambiguous. Protocols for documenting and categorizing triggering case histories
could result in more consistency among databases.

Identifying Which Soils Liquefied

The soil layer inferred to have liquefied in a triggering case history is commonly called the
“critical layer,” and the depth to this soil is called the “critical depth” (zcrit). The critical depth for
“no” cases is taken generally as that at the center of the soil layer judged to have the highest
liquefaction potential. The criteria used to select the critical layer in case studies, however, is
often not provided. This could result in inconsistencies among studies and uncertainty in
establishing a triggering relationship from the data. Early case history databases sometimes
contained multiple critical layers extracted from a single boring or sounding (e.g., Yegian and
Vitelli, 1983; Tokimatsu and Yoshimi, 1983; Seed et al., 1984; Stark and Olson, 1995). This
practice has been avoided in the development of more recent databases, as confirmed by study of
64 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

the databases by members of this study committee, in which the “weakest-link-in-the-chain”


concept is used to select a single critical layer in a soil profile (e.g., Cetin et al., 2000; Moss,
2003; Idriss and Boulanger, 2008; Kayen, et al., 2013: and Boulanger and Idriss, 2014).
In “yes” cases, the critical layer has sometimes been determined by comparing soil ejecta
from surficial liquefaction manifestations to samples obtained from borings (e.g., Liao and
Whitman, 1986; Green et al., 2011). In such cases, there is a possibility that the soil ejecta
include soil from the overlying strata. Water expelled from a liquefied soil, with or without
material from that soil itself, can entrain material from overlying strata as it flows from the
critical layer to the ground surface. For example, the 1989 Loma Prieta earthquake produced
boils of mud that erupted in an area where liquefiable sand underlies estuarine mud. Estuarine
mud may have liquefied, as suggested by Mejia (1998), or the mud may merely have been
fluidized by waters derived from the underlying sand, as suggested by Boulanger and colleagues
(1998). It is rarely simple to use trenches and borings to trace subsurface intrusions to the source
of the expelled water. Along the Columbia River, for instance, sediment peels from large
rectangular cores were used in an attempt to trace surficial dikes formed during the 1700
Cascadia earthquake downward to their roots (Takada and Atwater, 2004). The attempt was
confounded both by the apparent dissipation of subsurface dikes in cross-bedded sand and by the
likely presence of dikes and sills from earthquakes prior to the one responsible for the surficial
diking (see Figure 2 in Box 3.1).
General criteria for selecting the critical layer for paleoliquefaction as well as modern case
histories have been proposed in a few studies wherein consistency between the observed surface
manifestations and the depth, density, and thickness of the critical layer is emphasized (Green et
al., 2005; Olson et al., 2005; Green et al., 2014). Considerable judgment is still required to
implement these criteria, however, suggesting that differing interpretations of the same case
histories is likely (e.g., Boulanger and Idriss, 2014; Green et al., 2014). Use of multiple lines of
evidence to identify the critical layer and protocols for documenting and interpreting the critical
layer and its representative properties could reduce subjective interpretation and its associated
uncertainties.

Representative Properties of Soils That Liquefied

The soil in a triggering case history database is generally characterized by a single


normalized in situ test parameter (e.g., corrected SPT N-value). There is no standard protocol for
selecting this value where multiple in situ test parameter values are measured in the critical layer.
Some investigators have taken the smallest measured test parameter value as the representative
value of the critical layer (Liao and Whitman, 1986). Other investigators have used the average
value of the in situ test parameter within the critical layer (e.g., Cetin et al., 2000; Idriss and
Boulanger, 2008), in which case the selected thickness of the critical layer becomes important
because the measured test parameter values often vary with depth within a given stratum (e.g.,
Robertson, 2009; Boulanger and Idriss, 2014; Green et al., 2014).
Where averaging is used, the resulting representative test parameter can differ depending on
whether the measured test indices are averaged first and then corrections are applied, or whether
the corrections are applied to the measured values and the corrected values are then averaged
(Green et al., 2014). The differences between the two averaging approaches can be large for soils
with high fines contents. The committee is not aware of any justification for averaging before an
in situ resistance parameter is applied.
CASE HISTORIES 65

Pre- Versus Post-Earthquake Soil Properties

It has already been stated that most case histories in liquefaction triggering databases rely on
in situ test results obtained after liquefaction occurred. The inherent assumption in this
approach—namely, that characteristics of the critical layer do not change significantly as a result
of the earthquake—is supported to a limited extent by data from sites in California and New
Zealand, where both pre- and post-earthquake data are available (Holzer and Youd, 2007; Orense
et al., 2011). Some investigators, however, have reported large differences between pre- and
post-earthquake SPT N-values (Ohsaki, 1966). In those cases, soil layers with low pre-
earthquake N-values tended to densify (contract) while soils with high pre-earthquake N-values
tended to loosen (dilate).
It is difficult to reconcile the effects of an earthquake on the critical layer penetration
resistance. Therefore, liquefaction databases need to report the measured values of soil resistance
parameters and whether those measurements were made pre- or post-earthquake. Individual
investigators may then decide whether or not to make a correction to post-earthquake values (and
document any such correction) when interpreting a case history.

Identifying the Seismic Demand Imposed on the Critical Layer

In the absence of a downhole accelerometer array at the site, a detailed site-specific site
response analysis generally is the most accurate means to estimate the seismic demand imposed
on the critical layer (e.g., the shear stresses induced in the critical layer). Such analysis, however,
is possible only if the soil profile is well characterized down to bedrock, the Vs of the bedrock is
known, and representative bedrock input motions are available.
In lieu of detailed site response analyses, simplified procedures often are used to estimate
the seismic demand imposed on the critical layer. Several potential sources of ground motion
intensity data of varying degrees of uncertainty are used to estimate the intensity data for the case
histories. These include on-site ground surface motion recordings; near-site ground surface
motion recordings (preferably at sites with known Vs profiles) with intra-event residuals for
conditional PGA (Bradley, 2013) estimation; ShakeMap intensity measure values;1 event-
specific ground motion predictive equation (GMPE) values; values back-estimated from
observed structural performance (e.g., overturning of headstones); and region-specific as well as
general GMPE values.
Case history databases need to include enough information, such as the magnitude, source
mechanism, and site-to-source distance of the earthquake, to permit independent assessment of
demand value by the user. When the database includes a demand parameter (e.g., the PGA at the
ground surface or the CSR at the critical depth), the method used to establish this parameter
should be documented.

Limitations of Case History Database Coverage

While there are hundreds of liquefaction triggering case histories (see Table 3.2), only a few
dozen plot near the curves that separate zones of liquefaction and no liquefaction: that is,
1
E.g., http://earthquake.usgs.gov/research/shakemap.
66 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

deterministic plots of CRR versus soil liquefaction resistance parameter (N1)60cs, Qc1Ncs, or Vs1
(see Chapter 4). This limited number of influential case histories increases the uncertainty in the
location of the CRR curve. In addition to the limited case histories near the deterministic CRR
boundary line between “yes” and “no” cases, there is also a paucity of data for depths of less
than 2 meters and greater than 15 meters, for soils with nonplastic fines content greater than 35%
by weight, for gravelly soils, and at sloping ground sites.
A maximum depth at which liquefaction can be triggered is an important, yet controversial,
factor when evaluating the risk to projects from liquefaction. As may be observed from Table
3.2, the plots shown in Figure 3.1, and in Appendix C, the maximum depth to the center of the
critical layer listed in recently compiled case history databases is 20.5 meters, but some
researchers place the maximum value for that same case history at considerably shallower
depth.2 The next deepest cited cases for liquefaction are approximately 18 meters for two sites
(Kayen et al., 2013), but it is uncertain whether these two sites are included in the other
databases. The maximum depth to the center of the critical layer is approximately 15 meters in
most databases. The absence of “yes” case histories with a critical layer depth of greater than 15
to 20 meters does not mean liquefaction cannot be triggered at such depths. It may just indicate
that surface manifestations of liquefaction are unlikely for level ground conditions (i.e., where
the initial static shear stress has little to no influence on the triggering or manifestation of
liquefaction) if liquefaction occurs at those or greater depths. For example, it is believed that
soils liquefied at depths up to 25 meters or greater in the Lower San Fernando Dam during the
1971 San Fernando earthquake (Castro et al., 1992), resulting in the upstream slide of the dam. It
is unknown whether surficial manifestations of liquefaction would have been evident in the
absence of the upstream slide, or whether the soil at this depth would have liquefied in the
absence of the high initial static shear stress conditions. Data from instrumented sites where
strong ground shaking is expected and liquefiable layers at depths of greater than 15 meters are
present could indicate that liquefaction can be triggered at depth. Additionally, centrifuge tests
may provide insights into the maximum depth at which liquefaction can occur.
Limitations of the case history databases are expected to decrease given the growing size
and quality of field case history databases, especially when data from the 2010-2011 earthquakes
in Chile, Japan, and New Zealand and other recent earthquakes3 are included. Emphasis is
needed on documenting quality case histories with parameter ranges beyond those of case
histories currently in the database. Case histories that approach and extend the bounds of the data
with respect to grain size (i.e., low plasticity silts and gravelly soils), depth (i.e., less than about 2
meters and greater than about 15 meters), earthquake magnitude (i.e., greater than magnitude
7.8), and from sloping ground sites are of particular interest. Continued collection of case history
data within the ranges covered by existing databases is also warranted to reduce associated
uncertainties.

2
Cetin and colleagues (2004) interpreted the critical depth to liquefaction at the Kushiro Port Seismograph Station
site during the 1993 Kushiro-Oki, Japan, earthquake to be 20.5 m, but Kayen and colleagues (2013) and Boulanger
and Idriss (2014) reinterpreted the critical depth for this case history as 4.5 m and 3.8 m, respectively.
3
Including the 1999 events in Turkey and Taiwan and the 2004 and 2007 events in western Japan.
CASE HISTORIES 67

Balance Between Liquefied and Non-Liquefied Sites

Imbalance in the number of “yes” and “no” cases can be a significant source of bias when
establishing CRR curves. While recently compiled SPT databases do balance the numbers of
“yes” and “no” cases fairly well, recently compiled CPT and Vs databases contain significantly
more “yes” than “no” cases (see Table 3.2). Information from profiles where liquefaction is
known to have occurred—for example, where surface manifestations of liquefaction are
observed—often attracts more attention than that from nearby areas where no surficial evidence
of liquefaction is observed. As a result, data are not collected for those nearby profiles that may
have liquefied without surface manifestations; that may be susceptible to liquefaction but did not
liquefy under the specific earthquake loads; or that are not liquefiable. To minimize bias, the
numbers of “yes” and “no” cases need to be balanced when developing liquefaction resistance
correlations (e.g., CRR curves).

Minimum Earthquake Magnitude

Although M 5.9 is the smallest earthquake magnitude represented in Table 3.2 and Appendix
C, sand boils in the dry lake bed of Soda Lake, a quarry-tailing deposit, were observed following
an M 4.6 aftershock of the 1989 Loma Prieta earthquake (Sims and Garvin, 1995)—the
minimum magnitude earthquake reported to have triggered liquefaction. In this case, ejecta were
vented through existing dikes formed during the main shock and an earlier aftershock. It is
possible that no surface manifestations would have occurred if ejecta pathways had not already
existed. The lowest magnitude main shock event documented to have triggered liquefaction is an
M of about 5 (Ambraseys, 1988).
The absence of case histories for earthquakes with an M less than 5.9 in the databases
represents a significant shortcoming because the potential for surface manifestations and damage
to fragile structures due to liquefaction in events having magnitudes less than about 5.9 cannot be
completely discounted. Developers of liquefaction triggering analyses typically provide
magnitude scaling factors for earthquakes as small as M 5.5. The variability among the M 5.5
magnitude scaling factors proposed by various investigators (discussed in Chapter 4) is greater
than the variability at larger magnitudes and is a major source of uncertainty in liquefaction
analysis. Because magnitude scaling factors often are extrapolated in practice to magnitudes
even smaller than 5.5, the uncertainty is exacerbated. Therefore, case histories of the
performance of liquefiable soil deposits subject to earthquakes of magnitudes 5.5 or smaller,
whether or not surface manifestations of liquefaction or liquefaction-induced damage are
observed, are valuable, if not essential, for reducing the uncertainty in predicted liquefaction
hazard due to small magnitude events.

Predominance of Plate-Boundary Settings

Shallow ruptures on faults along or near tectonic plate boundaries, as may occur in
California, and shallow intra-slab events associated with subduction zones, as may occur in
Japan, are the source mechanisms represented most often in the liquefaction case history
databases. Comparatively few case histories are available for great earthquakes—that is, of
magnitude 8 to 9—on the plate-boundary thrust zones. But as they become available, case
68 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

histories from the 2010 Maule, Chile, and 2011 Tohoku, Japan, earthquakes will provide more
examples from subduction zones.
Also in short supply are well-documented case histories of liquefaction from earthquakes in
stable continental regions (e.g., the New Madrid and Charleston seismic zones in the United
States). Because the bedrock is denser, colder, and less fractured in these regions than in near
plate boundaries, bedrock earthquake motions are expected to be higher in amplitude, shorter in
duration, and richer in higher frequencies than in shallow crustal or subduction zone earthquakes.
It has been suggested that these differences in ground motions may lead to difference in
liquefaction triggering behavior (Semple, 2013). These interior earthquakes happen so rarely,
however, that few cases histories from this tectonic setting are likely to emerge in the coming
decades. This predicament adds to the importance of laboratory and analytical studies of the
impact on liquefaction behavior of the nature of ground motions in these regions.

Influence of Soil Age on Liquefaction Potential

The vast majority of case histories in the liquefaction triggering databases represent either
man-made fills or Holocene alluvial and fluvial sediments (Robertson and Wride, 1998; Youd et
al., 2001). Based on geologic evidence, Youd and Hoose (1977) and Youd and Perkins (1978)
postulated a generalized ranking of the liquefaction susceptibility of sedimentary deposits by age
and sedimentary environment. In this ranking, sedimentary deposits become less susceptible to
liquefaction as they age. Deposits less than 500 years old are generally ranked most susceptible,
and deposits more than about 2 million years old are ranked least susceptible. Revisiting case
studies containing Holocene soils and fills may provide beneficial information on the rate of
strength gain with age. Most deposits of Pleistocene age (between about 10,000 and 2 million
years) are ranked low or very low in liquefaction susceptibility, but Pleistocene deposits are
known to have liquefied in historic earthquakes. Alluvial fan deposits estimated to be 10,000-
15,000 years old (Andrus, 1986) liquefied during the October 28, 1983 Borah Peak, Idaho,
earthquake (M 6.9; Andrus and Youd, 1987). Far older Pleistocene deposits liquefied during the
September 1, 1886 Charleston, South Carolina, earthquake (M 7.3; Martin and Clough, 1994).
The Charleston area contains geologic evidence for repeated liquefaction at intervals
averaging 500-600 years over the past 6,000 years (Talwani and Schaeffer, 2001). As a result,
even though these deposits are relatively old from a depositional point of view, from a
geotechnical perspective (i.e., time since last significant disturbance), they are much younger
(see, e.g., Andrus et al., 2009). The Charleston data suggest that liquefaction will “reset the
clock” with respect to the age of a soil deposit for a liquefaction assessment. Nevertheless, it is
not clear if “near liquefaction” (i.e., significant pore-pressure generation without liquefaction)
may also reset the clock, and studies are therefore warranted on this point.

Case History Quality Rating

Cetin and colleagues (2000) made a major advance in case history compilation in their SPT
database, wherein procedures for quantifying uncertainties and assigning an overall quality rating
for case histories were proposed. They define four case history quality classes, and the criteria
for each relates to the quality and thoroughness of subsurface characterization and the
quantification of the seismic demand—information needed to interpret a case history properly.
The quality ratings allow the model developers to include, exclude, or weight cases based on the
CASE HISTORIES 69

certainty in their interpretation. A quality rating system of the type proposed by Cetin and
colleagues (2000) would enhance the value of liquefaction case history databases.

SITE-SPECIFIC GROUND MOTION RECORDINGS

Installation of strong motion instruments has increased in the last two decades and has
resulted in ground motion recordings at sites underlain by liquefied soil. Recorded motions at
those sites reflect the occurrence of liquefaction and can be used to better understand liquefaction
triggering and the response of liquefied profiles. Most of the sites were equipped with only
ground surface instruments, but a few of those locations also include downhole instruments in
vertical arrays.

Vertical Arrays

Vertical ground motion instrument arrays, in which strong motion instruments are placed at
the ground surface and at one or more depths in nearby boreholes, can provide excellent
opportunities to understand the seismic response of soil profiles and of individual soil layers
between pairs of instruments at adjacent depths.4 They also provide data on the ground shaking
intensity at the time liquefaction is triggered. Vertical arrays (see Box 3.2) installed specifically
to better understand liquefaction behavior are still relatively rare, but such arrays may include
pore-pressure transducers and slope inclinometers to provide information on liquefaction
behavior, reducing uncertainties on the location of the critical layer and the intensity of ground
motions necessary to induce liquefaction. Shape-acceleration arrays (SAAs) (Danisch et al.,
2004; Zeghal et al., 2007) offer a flexible and economical means of measuring soil response in
vertical arrays. Installation of additional vertical arrays at sites with a high potential for
liquefaction would provide valuable case history data and detailed ground motion data for
calibration of numerical models when one or more of these vertical array sites actually liquefied.

4
The records obtained from such arrays provide data for development, calibration, and validation of empirical, semi-
empirical, and mechanistic models of liquefaction triggering.
70 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

BOX 3.2
Liquefaction Arrays

An example of a ground motion instrument array installed to study liquefaction is the Wildlife
Liquefaction Array (WLA) in southern California. Installed in 1982, the WLA consists of a surface and
downhole seismographs and five pore-pressure transducers (Holzer and Youd, 2007). In 1987, the WLA
was excited by the November 24, 1987 Elmore Ranch (M 6.0) earthquake, which did not trigger
liquefaction at the site, and the next day by the Superstition Hills (M 6.5) earthquake, which did trigger
liquefaction. The strong motion instruments recorded ground motions at the surface and immediately
below a 3- to 4-meter thick layer of loose, saturated silty sand, which liquefied. Of the five pore-pressure
transducers, four were determined to have malfunctioned, and the accuracy of the measurements of the
fifth is controversial (Hushmand et al., 1991; 1992; Holzer and Youd, 2007). Analyses of seismograph
records by Zeghal and Elgamal (1994) provided some of the first direct evidence of liquefaction- induced
phase transformation (i.e., from solid to liquid state). The WLA site was re-instrumented in 2001 and
subjected to numerous earthquakes of variable strength since that time. A nearby M 4.9 event that was
part of the August 2012 Brawley swarm produced peak acceleration greater than 0.3g and a peak pore-
pressure ratio of approximately 0.6 (Hauksson et al., 2013).
Another vertical array on Port Island, Japan, extended through soils known to have liquefied in the
1995 Hyogo-ken Nanbu (Kobe) earthquake (Sitar, 1995; Zeghal et al., 1996). The Port Island array did
not include pore-pressure transducers, but it did have strong motion instruments on the ground surface
and at depths of 16, 32, and 83 meters. Thick deposits of ejecta erupted near the array during the
earthquake. A number of investigators have used the Port Island array recordings to improve and validate
various effective stress site response models for pore-pressure development and liquefaction
(Cubrinovski et al., 1996; Elgamal et al., 1996).
Other vertical arrays with both accelerometers and pore-pressure transducers include the Lotung
array in Taiwan (Abrahamson et al., 1987), the Garner Valley array in southern California (Archuleta et
al., 1992), the CORSSA array in Greece (Pitilakis et al., 2004), the Llolleo array in Chile (Verdugo, 2009),
the Belleplaine array in the French Lesser Antilles (Gueguen et al., 2011), and the Seattle liquefaction
a
array.
________
a
See http://nees.ucsb.edu/facilities/seattle-liquefaction-array.

LATERAL SPREADING CASE HISTORIES

Compared to case history databases for liquefaction triggering, few lateral spreading case
history databases have been assembled (see, e.g., Hamada et al., 1986; Bartlett and Youd, 1995;
Rauch and Martin, 2000; and Youd, et al., 2002). Cases that are available have been
characterized almost exclusively using SPT. The database of 484 case histories compiled by
Youd and colleagues (2002), a revision of the Bartlett and Youd (1995) database, is among the
most recently published. The majority of the cases in this database are from the 1964 Niigata,
Japan, and the 1983 Nihonkai-Chubu, Japan, earthquakes. As is true for empirical relationships
to predict liquefaction triggering, the validity of empirical relationships for predicting the
intensity of lateral spread displacements is limited by the parameter ranges of the case histories
from which the relationships are derived. The geometric parameters used to characterize lateral
spreads by Youd and colleagues (2002) are defined in Figure 3.2, and the ranges of all of the
parameters used in the Youd and colleagues (2002) case histories are listed in Table 3.3.
Sources of bias in lateral spread case history databases are similar to those in liquefaction
triggering databases (e.g., tectonic setting, geologic age and setting, pre- versus post-earthquake
characterization of soil profiles, interdependence of case histories), but they are more significant
in some regards. For example, multiple displacement vectors at a lateral spread site are treated as
CASE HISTORIES 71

separate case histories by Bartlett and Youd (1995) and Youd and colleagues (2002). Rauch and
Martin (2000) modified the Bartlett and Youd (1995) database by grouping vectors from the
same spreading area, oriented in the same general direction, into a single case history. Using this
approach, the 467 cases listed in the Bartlett and Youd (1995) database were reduced to
approximately 70 case histories. Some case histories in the Rauch and Martin (2000) database,
however, were from adjacent sites and therefore may not be completely independent.
Consequently, further efforts to account for case history interdependence are needed when
developing lateral spread predictive relationships, such as using mixed effects regression
techniques (Pinheiro and Bates, 2000).
Lateral spread case history databases have not been subject to the same level of scrutiny as
have liquefaction triggering databases. Additionally, the lateral spreading phenomenon is
complex, and the resulting spatial spreading deformations and patterns are often characterized
simply (e.g., as a single lateral displacement). Often only a few in situ field tests are used to
characterize the soils across a lateral spread. This combination of factors results in fewer high-
quality, well-documented case histories. Protocols are needed for documenting and interpreting
lateral spreading case histories, particularly to address the issue of case history interdependence.
Formal quality rating schemes for lateral spread case histories need to be developed and clearly
stated by model developers that use the case histories. These ratings need to be taken into
consideration when developing, calibrating, or validating lateral spreading models.
Remote sensing technologies such as terrestrial and airborne Light Detection and Ranging
(LiDAR) and satellite imagery are being employed more commonly in post-earthquake
reconnaissance (e.g., Bray and Frost, 2010). Such technologies allow for more comprehensive
documentation of the displacements associated with lateral spreading. For example, optical
image correlation (Leprince et al., 2007) can be applied to pre- and post-earthquake satellite
imagery to develop lateral displacement measurements at high spatial resolution (e.g., Martin
and Rathje, 2014) as shown in Figure 3.3. Pre- and post-earthquake LiDAR-derived digital
elevation models can be used to compute settlement and lateral displacements associated with
liquefaction and lateral spreading (e.g., van Ballegooy et al., 2014b). Pre-earthquake data are
required, however, to use these approaches, and investing in baseline data collection in areas of
interest (e.g., Holocene deposits in seismically active areas) in anticipation of earthquakes would
make use of such methods possible. Additionally, such new approaches and technologies as
digital elevation models derived from digital photogrammetry and real-time processing of
LiDAR data collected from unmanned aerial vehicles are being explored for documenting and
interpreting case histories (Rathje and Franke, 2015).
72 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 3.2 Parameters used to characterize lateral spread case histories.


NOTE: DH = horizontal ground displacement; H = height of the free-face; L = horizontal distance from the freeface;
S = ground slope; and free-face ratio W = H/L. SOURCES: Bardet, J.P., T. Tobita, N. Mace, and J. Hu. 2002.
Regional modeling of liquefaction-induced ground deformation. Earthquake Spectra 18(1):19–46.; after Bartlett and
Youd, 1992. With permission from the Earthquake Engineering Research Institute.

TABLE 3.3 Parameter Ranges for Case History Database


Parameter Range
Depth of range (M; m) 6.4-9.2

Site-to-Source Distance (R; km) 0.2-100


a
Peak ground acceleration (g) 0.19-0.55

Cumulative thickness (m) of liquefiable soils in upper 0.01-19.7


20 m of soil (N1,60 < 15 blows/30 cm; T15)

Average fines content of soil comprising T 15 (D50; 0-70


mm)

Median particle size of soil comprising T 15 (D50; mm) 0.036-12

Slope angle from horizontal (S; degrees) 0.05-11

Ratio of height of free face to distance from free face 1-56.8


(W; %: see Figure 3.2)

Permanent horizontal deformation observed at the 0.01 -10.16


point of interest (DH; m)
a
PGAs from Bardet et al. (2002).
NOTE: Abbreviations used in Figure 3.2 provided.
SOURCE: Youd et al. (2002).
CASE HISTORIES 73

FIGURE 3.3 Amplitudes of horizontal displacement from optical image correlation displayed along with
observed crack locations from field reconnaissance of the 2011 Christchurch, New Zealand, earthquake.
This is an example of what can be derived from satellite imagery in situations where both pre- and post-
earthquake imagery is available at high resolution. SOURCE: Martin, J.G., and E.M. Rathje. 2014. Lateral
spread deformations from the 2010-2011 New Zealand earthquakes measured from satellite images and
optical image correlation. In Proceedings of the 10th U.S. National Conference in Earthquake
Engineering, 21–25 July, Anchorage, Alaska. Oakland, CA: Earthquake Engineering Research Institute.
With permission from the Earthquake Engineering Research Institute.

POST-LIQUEFACTION SHEAR STRENGTH CASE HISTORIES

Compilations of field case histories for post-liquefaction residual shear strength (i.e., large
deformation) of granular soils are significantly smaller and contain larger uncertainties than are
those for liquefaction triggering and lateral spreading. One of the first, if not the first, field case
history databases for residual shear strength consisted of just 12 case histories and included
earthquake-induced failures of several dams, dikes, embankments, natural slopes and lateral
spreads, an earthquake-induced bearing capacity failure of a four-story apartment building, and
the failure of two hydraulic-fill dams under construction (Seed, 1987). Normalized and fines-
corrected SPT N-values were provided for each case, as were estimates of the residual strength
from back-analyses based on initial and final geometries of the sliding mass system.
Subsequent researchers have added, reinterpreted, and filtered case histories based on
deformation mode, and they have applied quality control ratings to residual shear strength case
histories (e.g., Seed and Harder, 1990; Stark and Mesri, 1992; Olson and Stark, 2002; Idriss and
Boulanger, 2007; Robertson, 2010; and Kramer and Wang, 2015). Table 3.4 summarizes the
ranges of the variables in the case history databases for the most commonly used correlations. As
may be seen from this table, there are no case histories that have a corrected SPT N-value greater
than 15 blows/30 cm or a corrected CPT tip resistance greater than 85 atm. The validity of
empirical relationships for predicting residual strengths is limited by the parameter ranges of the
case histories from which the relationships are derived.
Evolution of the databases has not been linear: some databases exclude case studies that are
included in subsequent databases. For example, Idriss and Boulanger (2007) explicitly removed
lateral spread case histories from their residual shear strength database. Olson and Johnson
74 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

(2008), on the other hand, stated that their analyses showed that residual shear strengths back-
calculated from lateral spread cases were in good agreement with those from the Olson and Stark
(2002) residual shear strength case history database. Additionally, Stark and Mesri (1992), Olson
and Stark (2002), and Kramer and Wang (2015) express the residual shear strength as the
strength ratio (i.e., ratio of residual shear strength to the initial vertical effective stress), while the
other databases do not. The residual shear strength ratio of Kramer and Wang (2015) is a
function of the magnitude of the vertical effective stress. Also, Olson and Stark (2002) and
Kramer and Wang (2015) state that the strength ratios correlate just as well or better to
penetration resistances that are not corrected for the influence of fines content, while other
databases use corrected penetration resistances. Seed and Harder (1990), Olson and Stark (2002),
Idriss and Boulanger (2007), and Kramer and Wang (2015) considered dynamic (or kinematic)
effects in estimating the residual shear strength. Variability among databases complicates data
collection and case history interpretation. Accordingly, protocols need to be established for
documenting and interpreting post-liquefaction residual shear strength case histories.
The quality and reliability of case histories can range, and quality rating of case histories has
varied in databases in which they have been used. For example, Idriss and Boulanger (2007)
included only those cases with enough information to allow dynamic (or kinematic) effects to be
incorporated into the calculation of the residual shear strength. Robertson (2010) rated the case
histories from Class A to E based on details of the in situ characterization of soil properties.
Class A cases were characterized by the modern electronic CPT prior to failure, and subsequent
classes have less reliable characterizations (e.g., estimated soil properties based on judgment, the
SPT, mechanical CPT). Kramer and Wang (2015) applied a weighting factor of 1.0 to thoroughly
investigated and well-documented case histories and lower weighting factors (between 0.8 and
0.2) to other case histories. Applying a rating system to case histories in a database to be used for
establishing a post-liquefaction shear strength correlation is a welcome improvement. Because
all quality rating systems are inherently subjective, the rationale for the rating system and the
weighting factors used within it need to be documented in such a way that allows potential users
of the database and resulting regressions to be fully informed about the system employed.
CASE HISTORIES 75

TABLE 3.4 Variable Ranges in Post-Liquefaction Residual Strength Case History Databases
Seed
and Stark and Idriss and Kramer
Seed Harder Mesri Olson and Boulanger Robertson and Wang
a a
(1987) (1990) (1992) Stark (2002) (2007) (2010) (2015)
Number
of cases 12 17 20 33 18 34 31

N1,60-cs
(blows/30
3-15 3-15 3-15 0-11.5 3-15 - 1.1-12.6
cm)

qc1-cs
(atm) - - - 0-60 25-85 34-77 -

Fines
content 3-80 3-80 0-80 0-100 0-85 0-100 0-100
(%)
a
The listed penetration values for Seed (1987), Seed and Harder (1990), Stark and Mesri (1992), Idriss and
Boulanger (2007), and Robertson (2010) include fines content corrections (i.e., N 1,60-cs or qc1-cs). The listed
penetration values for Olson and Stark (2002) and Kramer and Wang (2015) do not include fines content corrections
(i.e., N1,60 or qc1).

NONTRADITIONAL SOURCES OF DATA

Recent developments in instrumentation and sensing, imaging technologies, and data


acquisition, processing, and display may provide new and unique opportunities to collect and
record data during and after strong earthquake shaking. For example, mobile phones with video
and Global Positioning System capabilities allow liquefaction features to be captured on video
and accurately located. While engineers often think of “response” as detailed measurements of
displacements, velocities, accelerations, porewater pressures, and other physical reactions, videos
of earthquake response by laypersons are more commonly available, and crowdsourcing of
earthquake response data is becoming possible. Coupled with detailed post-event deformation
measurements, crowdsourced data such as videos can aid the understanding of a case history.
While most efforts to date have focused on emergency response or humanitarian actions
(Starbird and Palen, 2013), opportunities to extract technical information such as locations of
liquefaction features and lateral spreading deformations also exist. Nontraditional types and
sources of data for documenting and interpreting liquefaction case histories may be a means to
expand existing databases.

ENHANCING DATABASE DEVELOPMENT

This chapter describes existing liquefaction case history databases and their limitations. The
impact of these limitations could be reduced if future case histories meet heightened and
consistent standards for documentation. Strong, community-based collaboration will be
necessary to establish standards to allow a more holistic understanding of liquefaction and its
consequences. Table 3.5 is an example draft checklist adapted from the Next-Generation
Liquefaction (NGL) project (see Box 3.3) for documenting liquefaction and related phenomena
76 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

case histories. This table represents a good starting point for a standard case history
documentation protocol, but it lacks certain important information such as the availability of
remote sensing data and whether data quality was rated.
In order for the NGL database to become the model for documenting case histories, the final
draft needs to be thoroughly vetted by the geotechnical community. When establishing a model
for documenting case histories, however, it should be noted that there are established
geotechnical data collection protocols (e.g., those by the Geotechnical Extreme Event
Reconnaissance organization)5 that need to be considered. Although it is unlikely that all of the
items described in a final protocol will be available for any given case history, providing as
much of this information as possible could increase the value and quality of the case histories.
Factors necessary to establish the quality of the case history (if not a quality rating system) in the
database are important. Engineering practice will improve when case history documentation
consistently includes factors that describe the quality of the data, and when case history quality
ratings are considered during development, calibration, and validation of post-liquefaction
residual shear strength models. Development of the quality rating can be left to individual
investigators developing correlations, provided their case history documentation includes those
factors necessary to develop a rating system.

5
See http://www.geerassociation.org/reconnaissance.htm.
CASE HISTORIES 77

TABLE 3.5 Example Checklist for Documenting Liquefaction and Related Phenomena Case Histories
Data Type Desired Documentation Comments
Site overview Longitude and Latitude of Site

Topography, Ground Slope Proximity to topographic irregularities

Presence/Proximity of Building structures, buried infrastructure,


Structures embankments, etc.

Geologic Natural or Man-Made Fill Construction records for fills (dates, methods of
setting and site placement)
geology
Depositional Environments

Age of Deposits

Spatial Variability Both lateral and vertical variability, potential for


layering (and void redistribution)

Evidence of Description of Sand Boils Severity of liquefaction/dimensions of boil/amount of


Liquefaction ejecta, color/classification of ejecta, etc.

Ground Movement Ground crack widths, lengths, orientations; LiDAR


images, remote sensing measurements

Ground Settlement Amount and method of measurement; reference


point for measurements

Ground Motion Characteristics E.g., dilational spikes in acceleration time history

Damage to Structures and Post-triggering consolidation settlement, bearing


Other Infrastructure capacity failure, broken buried pipelines, etc.

Visual Accounts Photographs of before and after, videos from


observers or security cameras, interviews with eye
witnesses, etc. For lateral spreading and residual
shear strength case histories, LiDAR scans and
high-resolution aerial photographs for both pre- and
post-event or topographic surveys.
Ground Motions Evidence of rapid changes in frequency content,
dilation-induced acceleration spikes

Site Borings Number, locations relative to liquefaction features,


Characterization drilling data (i.e., rotary wash, hollow stem, casing)

SPT System Hammer type, rod type and length, drill bit, lifting
system, sampler diameter (inside and outside),
energy measurements

SPT Data 6/6/6 blow counts, soil description, sample recovery

CPT Soundings Number, locations relative to liquefaction feature,


availability of CPT samples
78 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Data Type Desired Documentation Comments


Vs Data Type of measurement (downhole, seismic CPT,
spectral analysis of surface waves, multichannel
analysis of surface waves, etc.); locations

Vp Data For determination of depth to saturation

Groundwater Data Depth at time of field testing and at time of


earthquake, daily and seasonal fluctuations

Laboratory Data Grain Size Distribution Fines content

Atterberg Limits PI, LL

Soil Mineralogy Silicate, feldspar, calcareous, etc.

Angularity of Particles
Water Content, Unit Weight
Tests of Mechanical Behavior Consolidation, undrained strength, cyclic tests, etc.

Ground Motions Recorded Motions PGA, acceleration time history; location of nearby
strong motion stations; type of instrument/housing;
links to recorded motions

Source/Path/Site Parameters Moment magnitude, source-to-site distance (Rjb,


Rrup, etc.), style of faulting, Vs30, presence of basin,
basin depth, hanging wall, etc.; event type (shallow
crustal in active region, subduction, stable
continental region)

ShakeMaps Event-specific

Ground Shaking Intensity Structural effects or damage caused by ground


vibratory motion (e.g., overturning and sliding of
rigid objects or yielding of moment frames)

Ground Motion Prediction Regionally applicable where available; event terms,


Equations mapped within-event residuals

SOURCE: Personal communication, S. Kramer.


CASE HISTORIES 79

BOX 3.3
Next-Generation Liquefaction (NGL) Project

The development of a community database of liquefaction-related phenomena case histories has


been initiated by the Pacific Earthquake Engineering Research (PEER) Center. PEER’s Next-Generation
Liquefaction (NGL) project involves a partnership between PEER and various public agencies in the
United States, Japan, New Zealand, and Taiwan (Stewart et al., 2015). NGL researchers will assemble
and document existing case history data and obtain and document new case history data. Case histories
for triggering, lateral spreading, and settlement will be included. The NGL documentation effort will also
include laboratory, physical model, field, and numerical studies on key aspects of liquefaction triggering
and related phenomena that are poorly constrained by the current field case history database.
Experienced liquefaction researchers will vet the database as it is populated.
Recognizing that multiple viable technical interpretations of the data by knowledgeable researchers
are possible, independent teams will be assembled to develop predictive models based on the community
database. The entire process of database and model development will be undertaken with regular
communication among investigators via project coordination meetings and with public workshops to
enable community engagement and input. Predictive model developers will be required to explicitly justify
the exclusion of data from their models. A major advantage of this approach is that the resulting model
predictions will reflect the genuine epistemic (modeling) uncertainty associated with alternate methods of
data interpretation of a common data set.
As of this writing, the NGL project has held five workshops with U.S. and international partners since
2013. Initial projects are under way to collect supplemental field data at strategically targeted case history
sites in Japan and New Zealand.
4
The Simplified Stress-Based Approach to
Triggering Assessment

Key Findings and Conclusions

 Only the specific adjustment factors used for development of a particular liquefaction triggering
relationship are suitable for use with that relationship.
 Extrapolations of liquefaction triggering relationships outside the conditions supported by case
histories need to be based on laboratory data and soil mechanics principles.
 Liquefaction triggering analysis needs to move toward a fully probabilistic approach that includes
uncertainties in the earthquake shaking, the liquefaction resistance of the soil, and the different
empirical liquefaction triggering models.
 Recommendations from recent studies related to the magnitude scaling factor need to be vetted by
the profession for their practical application. This will also be true for any updates and revisions based
on new data from large magnitude earthquakes.
 Empirical methods used for assessing liquefaction triggering have their merits and have served the
technical community well, but each is flawed. Practice could be improved by applying multiple
methods, assessing any source of discrepancy among results, and using professional judgment to
resolve those discrepancies.

Liquefaction is triggered when the generation of large porewater pressures forces a soil to
lose significant stiffness and strength and, under some circumstances, to behave like a viscous
fluid (see Chapter 2). A first step in a liquefaction evaluation is to assess whether the expected
seismic shaking is large enough to trigger liquefaction. The evaluation usually includes site
characterization: that is, studying the geologic, hydrologic, and geotechnical conditions that
influence the liquefaction potential of a site. The quantitative component of the evaluation is
usually calibrated to the shaking levels and soil properties that have or have not resulted in
liquefaction during previous earthquakes (see Chapter 3). The consequences associated with
liquefaction triggering are assessed separately (see Chapters 6 and 7).
The method known as the simplified stress-based procedure (Whitman, 1971; Seed and
Idriss, 1971; Seed et al., 1985) is the one most commonly used in practice to predict liquefaction

80
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 81

triggering (see Chapter 1, Box 1.3) and is the subject of this chapter. A factor of safety (FS)
against liquefaction triggering, defined as the ratio between the seismic loading required to
trigger liquefaction (i.e., the liquefaction resistance) and the seismic loading expected from the
earthquake (i.e., the seismic demand), is computed. Both the seismic demand and the
liquefaction resistance are characterized as cyclic stress ratios, defined as the ratio of the cyclic

shear stress (cyc) to the initial vertical effective stress (𝜎𝑣𝑜 ). The seismic demand is the
earthquake-induced cyclic stress ratio (CSR), and the liquefaction resistance is the cyclic
resistance ratio (CRR): that is, the cyclic stress ratio required to trigger liquefaction. The most
common approaches used in practice to compute the seismic demand (CSR) and liquefaction
resistance (CRR) are described in the next sections. Issues related to the collection of appropriate
geotechnical field data for a stress-based liquefaction triggering assessment are discussed. The
chapter ends with a discussion of the interpretation of simplified stress-based liquefaction
triggering analyses, including a discussion on the use of probabilistic methods.

SEISMIC DEMAND

The CSR represents the cyclic shear stress at a given depth due to earthquake shaking. At
sites where the ground slope is less than 6% (per Youd et al., 2001), the earthquake-induced
shear stresses are assumed to be due solely to the vertical propagation of shear waves and can be
computed by one-dimensional dynamic response analysis of the soil deposit (Seed and Idriss,
1971). Because dynamic response analyses were difficult and time-consuming to perform when
the method was developed, however, and because it was challenging to obtain the required site
characterization and to identify appropriate input ground motions, Seed and Idriss (1971)
proposed the following simplified equation (see Equation 4.1 below), based on Newton’s second
law, to compute a representative CSR for an earthquake magnitude M = m:

0.65 ∙ 𝜏𝑐𝑦𝑐,𝑚𝑎𝑥 (𝑧) 0.65 ∙ 𝑃𝐺𝐴/𝑔 ∙ 𝜎𝑣 (𝑧) ∙ 𝑟𝑑 (𝑧)


𝐶𝑆𝑅𝑴=𝑚 (𝑧) = ′ (𝑧)
= ′ (𝑧)
(4.1)
𝜎𝑣𝑜 𝜎𝑣𝑜

where cyc,max is the maximum cyclic shear stress at depth z, 𝜎′𝑣𝑜 is the initial vertical effective
stress at depth z, PGA is the horizontal peak ground acceleration at the ground surface, g is the
acceleration of gravity, 𝜎𝑣 is the total vertical stress, and 𝑟𝑑 is a depth-dependent shear stress
reduction coefficient that accounts for the nonrigid response of the soil deposit. The 0.65 factor
was introduced to reduce the CSR from the peak value of the earthquake-induced shear stress,
which occurs only once during the earthquake, to a more representative value that occurs multiple
times during strong shaking. The value of 𝑟𝑑 is equal to 1.0 at the ground surface and decreases
with depth below the surface to account for the nonrigid response of the soil column subjected to
the vertically propagating shear wave.
The cyclic stress ratio required to initiate liquefaction (i.e., the liquefaction resistance, CRR)
decreases with increasing number of cycles of loading (see Chapter 2); therefore, the seismic
loading must be associated with a number of loading cycles. Earthquake magnitude is used as a
proxy for the number of loading cycles because the duration of shaking and the
associated number of loading cycles correlate with earthquake magnitude. The CSR is
adjusted using a magnitude scaling factor (MSF) to compute an equivalent CSR for a
reference M = 7.5 (see Equation 4.2 below):
82 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

1 0.65 ∙ 𝑃𝐺𝐴/𝑔 ∙ 𝜎𝑣 (𝑧) ∙ 𝑟𝑑 (𝑧) 1


𝐶𝑆𝑅𝑀=7.5 (𝑧) = 𝐶𝑆𝑅𝑀=𝑚 (𝑧) ∙ = ′ (𝑧)
∙ (4.2)
𝑀𝑆𝐹 𝜎𝑣𝑜 𝑀𝑆𝐹

MSFs were first developed by Seed and Idriss (1982) from estimates of the typical number
of cycles of loading for different earthquake magnitudes and laboratory test results. Multiple
alternative MSFs, some of which are shown in Figure 4.1, have been proposed since then by
researchers using various techniques. The 1996/1998 National Center for Earthquake
Engineering Research (NCEER)/National Science Foundation (NSF) Workshops on liquefaction
evaluation (Youd et al., 2001) recommended a range of MSFs for use in practice (Youd et al.,
2001). More recently, Cetin and colleagues (2004), Cetin and Bilge (2012), and Boulanger and
Idriss (2014) provided revised recommendations for MSFs. Cetin and colleagues (2004) derived
their MSFs directly from the liquefaction case history database by including earthquake
magnitude in the regression analysis they used to establish the triggering relationship for
liquefaction resistance (described in the next section). But, because a large percentage of data in
current liquefaction databases comes from earthquakes of M between about 6.9 and 7.6 (see
Chapter 3 and Appendix C), where the MSF is relatively close to 1.0, statistical derivation of
MSFs outside of this magnitude range can be questionable. Cetin and Bilge (2012) developed an
MSF that depends on relative density and overburden stress. Boulanger and Idriss (2014) applied
a cap to the MSF for small magnitude earthquakes (M < 5.25) to account for the dominance of a
single large acceleration pulse in these small magnitude events and, based on work by Kishida
and Tsai (2014), they made the MSF dependent on a soil type parameter that described the slope
of the curve relating the CRR to the number of cycles to liquefaction. Thus, the MSF for dense
cohesionless soil is close to the “Idriss” curve shown in Figure 4.1. As the density (or penetration
resistance) decreases, however, the MSF curve flattens; this results in smaller values than Youd
and colleagues’ (2001) recommended range for M < 7.5 (thereby increasing the liquefaction
potential) and larger values for M > 7.5 (thereby decreasing the liquefaction potential).
Prior to 2001, the MSF depended only on earthquake magnitude. Revisions to MSFs since
then represent a major departure from this practice, but there is no consensus on which of the
more recent recommendations, if any, are appropriate for practical application. Furthermore, the
impacts of case history data on MSF from recent large magnitude earthquakes in Chile and Japan
have yet to be considered. MSF values also impact recommendations for the number of uniform
cycles of loading to use in laboratory-based evaluations of liquefaction triggering. MSFs and the
number of uniform cycles of loading for laboratory evaluation of liquefaction triggering, N eq, are
intimately related, as discussed in Chapter 5. As is also noted in Chapter 5, recent studies suggest
that Neq is negatively correlated with the peak ground acceleration (see, e.g., Biondi et al., 2004),
that Neq is a function of site-to-source distance (see, e.g., Liu, et al., 2001; Green and Terri, 2005;
Stafford and Bommer, 2009), and that Neq may vary with depth in the soil profile (see, e.g.,
Green and Terri, 2005). If Neq is dependent on these factors, then the MSF should also be
dependent on these same factors. Given the importance of the MSF when evaluating liquefaction
triggering for both small (e.g., M < 6.5) and large (e.g., M > 8.0) earthquakes, as well as its
importance to laboratory evaluation of liquefaction triggering, research to identify and resolve
inconsistencies related to the various MSF and Neq recommendations made since 2001 is
warranted. New case history data from recent large magnitude earthquakes need to be
incorporated into these recommendations.
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 83

The 0.65 factor found in Equations 4.1 and 4.2 was originally proposed as a way to relate
the number of loading cycles from an irregular earthquake loading to the number of loading
cycles from uniform cyclic loading. Although this value is somewhat arbitrary and was
unnecessary once MSFs were introduced, 0.65 is still the standard due to historical precedent.
Equations 4.1 and 4.2 require estimates of PGA, M, and 𝑟𝑑 to compute the CSR. Because
the required PGA is at the ground surface of a soil site, the PGA used in a liquefaction triggering
analysis must account for the effects of the near-surface soil conditions on ground shaking.
Traditionally, ground motion hazards were expressed as scenario events, so the PGA and M
values used in a liquefaction triggering assessment were established deterministically.
Probabilistic seismic hazard analysis (PSHA), however, has become the standard for defining
ground motions for seismic design for some applications. One challenge using a PSHA-derived
PGA is that there is no single earthquake magnitude associated with it from which to compute
the MSF. If the magnitude deaggregation1 data associated with the PSHA-derived PGA are
available, the associated FS against liquefaction triggering may be calculated by parsing the
deaggregated magnitude-distance pairs into magnitude bins; calculating the FS associated with
each magnitude bin using the associated MSF; calculating the contribution from each magnitude
bin to the design FS by multiplying the FS associated with the magnitude bin by the sum of the
relative contributions of each magnitude-distance pair in the bin; and summing up the
contribution to the design safety factors from all of the magnitude bins (Idriss, 1985, Finn, 2007).
If the deaggregation data are not available, engineering judgment may be required to select a
representative magnitude from the earthquake sources (e.g., active faults) associated with the
probabilistic PGA. This can be a relatively straightforward process if the seismic environment is
dominated by a single large magnitude earthquake source. When the seismic environment is
characterized by a number of sources of different magnitude, however, or by a fault capable of
generating earthquakes of varying magnitudes, identifying a single representative magnitude is a
challenging task that should be delegated to an experienced professional (e.g., an experienced
geotechnical earthquake engineer or engineering seismologist).
Different options are available for conducting a probabilistic liquefaction triggering
assessment using the stress-based method: these include a single-scenario option in which the
user can manually enter PGA and M; a multiple-scenario approach in which the deaggregation
data are integrated to obtain a marginal magnitude distribution, for each value of which the FS is
computed and then combined using the magnitude probabilities as weighting factors; and a
complete probabilistic liquefaction hazard assessment (PLHA) that considers the magnitude
distributions at all return periods. Other approaches have also been proposed, but it is essential to
recognize that the simplified, stress-based triggering assessment methods are empirical and
cannot be manipulated excessively without losing their reliability. A more comprehensive
probabilistic liquefaction assessment based on a PSHA takes into account the relative
contributions to the liquefaction hazard from the full range of ground motion intensities
associated with each magnitude-distance pair. This type of liquefaction assessment is discussed
in Chapter 9 within the context of an integrated reliability assessment and performance-based
seismic design.

1
Magnitude deaggregation refers to the process of dividing the family of earthquakes that contribute to the
probabilistic design ground motion into magnitude-distance pairs and calculating the relative contribution of each
pair to the design ground motion.
84 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

The depth-dependent shear stress reduction coefficient, 𝑟𝑑 , is a function of the nonrigid


response of the soil deposit (characterized in the small strain regime by the shear wave velocity
[Vs] profile at the site) as well as the characteristics of the earthquake waves traveling through
the soil. As a result, the range of possible 𝑟𝑑 values is large. Figure 4.2a illustrates the variability
in 𝑟𝑑 values computed for a range of site conditions. Seed and Idriss (1971) initially proposed a
relationship between 𝑟𝑑 and depth that they developed from a limited number of dynamic
response analyses for a range of generic site conditions. Using additional site response analyses,
Idriss (1999) modified the Seed and Idriss (1971) 𝑟𝑑 relationship to make 𝑟𝑑 depth and
magnitude dependent. The Idriss (1999) relationship is used to develop the triggering
relationships of Idriss and Boulanger (2008) and Boulanger and Idriss (2014).
More recently developed 𝑟𝑑 relationships (see, e.g., Cetin and Seed, 2004; Kishida et al.,
2009) take into account additional factors, including the average Vs. Incorporating additional
factors may produce more accurate estimates of 𝑟𝑑 , but the variability remains large. Figure 4.2b
shows the depth-dependent 𝑟𝑑 factors from Seed and Idriss (1971), Idriss (1999), and Cetin and
Seed (2004) for an average Vs of 150 m/s over the top 12 m, PGA = 0.4 g, and M = 7.5. The
Cetin and Seed (2004) relationship predicts significantly smaller values of 𝑟𝑑 than either of the
other relationships, though the differences between the relationships are smaller for stiffer soil
profiles. Differences in the Vs profiles analyzed by Cetin and Seed (2004) compared with those
used by either Seed and Idriss (1971) or Idriss (1999) are the main reasons for the differences in
𝑟𝑑 , although the different modulus reduction and damping curves used by these investigators also
influenced their 𝑟𝑑 values. Cetin and Seed (2004) performed site response analyses for 50 well-
characterized sites that previously liquefied in an earthquake, while the other studies analyzed
mostly hypothetical sites. As a result, the median values of 𝑟𝑑 predicted by the Cetin and Seed
(2004) relationship are considered more representative of the induced cyclic shear stress at
liquefaction sites than are other 𝑟𝑑 relationships. Because the position of data points on a
liquefaction triggering plot—and thus the position of the triggering curve itself—are intimately
related to the evaluation of 𝑟𝑑 , however, the 𝑟𝑑 relationship developed by Cetin and Seed (2004)
may not be appropriate for use with triggering relationships developed using the other 𝑟𝑑
relationships.
In place of the simplified expressions in Equations 4.1 and 4.2, the earthquake-induced
shear stress can be computed directly from site-specific dynamic response analyses (Kramer,
1996). The most common method used to conduct a site-specific dynamic response analysis is a
one-dimensional, total stress, equivalent-linear analysis. A one-dimensional nonlinear analysis, a
two-dimensional equivalent linear analysis, and two-dimensional nonlinear analysis are
sometimes used as well (Kramer, 1996). Many engineers prefer to use dynamic response analysis
to evaluate the CSR because the site-specific Vs profile and expected ground shaking at the site
are considered in the analysis. Note, however, that the use of site-specific dynamic response
analysis to assess the CSR for a simplified liquefaction analysis may not be consistent with the
empirical liquefaction curves used to establish liquefaction resistance if those curves are
developed using 𝑟𝑑 relationships that are biased relative to site response analysis (as discussed
later in this chapter). If the empirical 𝑟𝑑 relationship associated with the particular liquefaction
method is not an unbiased estimator of 𝑟𝑑 (e.g., for the triggering relationships developed by
Idriss and Boulanger [2008] and Boulanger and Idriss [2014]), the empirical 𝑟𝑑 relationship
associated with the particular liquefaction method should be used to compute induced shear
stresses at depth, and the results of site response analyses should be used only to refine the
estimate of the PGA at the ground surface. Furthermore, effective stress dynamic response
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 85

analysis with pore-pressure generation should never be used to estimate the seismic loading,
because the seismic loading for the simplified, stress-based procedure ignores the effects of pore-
pressure generation.
Estimating the shear stress from dynamic response analysis requires more information than
does the simplified equation: it requires the Vs distribution for the entire soil profile and the
underlying bedrock; the strain-dependent nonlinear soil properties (e.g., variation of shear
modulus and damping ratio with shear strain); and a suite of appropriate input earthquake
motions. The selected input motions must match the intensity and frequency content of shaking
expected at a site, and multiple time histories must be used to capture the variability inherent to
earthquake ground motions. For instance, the International Building Code (ICC, 2011), through
incorporation by reference of ASCE 7 (Chapter 21),2 recommends that at least five different
ground motions be used in site response analysis (ASCE, 2010). Alternatively, a random
vibration theory (RVT) approach to equivalent-linear site response analysis can be used (see,
e.g., Rathje and Ozbey, 2006; Kottke and Rathje, 2008). The RVT approach has the advantage of
not requiring the selection of input time series.
There is considerable uncertainty in the CSR expected at a site due to uncertainty and
variability in the site characterization, variability in the 𝑟𝑑 expressions (see, e.g., Figure 4.2), and
uncertainty in the ground motion PGA and associated earthquake magnitude. The uncertainty in
PGA is somewhat taken into account through the use of a PSHA to estimate the ground motion.
Accounting for the other sources of uncertainty may require a more complete PLHA (described
later in this chapter).

FIGURE 4.1 Recommended magnitude scaling factors (MSFs) from Youd and colleagues (1997a). MSFs derived by
various investigators prior to 1997 are also shown.

2
See http://ascelibrary.org/doi/book/10.1061/9780784412916.
86 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

(a) (b)
FIGURE 4.2 (a) Computed shear stress reduction coefficients (𝑟𝑑 ) for a range of site conditions and input motions.
The solid curves indicate the mean and standard deviations in calculated values of rd for different site conditions and
input motions (gray lines). (b) Proposed 𝑟𝑑 relationships from different researchers. SOURCE: (a) Cetin, K.O., and
R.B. Seed. 2004. Nonlinear shear mass participation factor (rd) for cyclic shear stress ratio evaluation. Soil Dynamics
and Earthquake Engineering 24(2):103–113. With permission from Elsevier. (b) Courtesy of E. Rathje.

LIQUEFACTION RESISTANCE

As already described, the resistance to liquefaction for a soil is expressed as the CRR, which
is the CSR required to trigger liquefaction. As discussed in Chapter 2, the in situ CRR cannot be
measured reliably in the laboratory for most liquefiable soils. Therefore, the simplified method
makes use of field observations to develop relationships between measured in situ parameters
and CRR. These relationships are derived from case histories of liquefaction and no liquefaction
from many sites and many earthquakes (see Chapter 3). The most common in situ parameters
used for this purpose (see Chapter 2) are the standard penetration test (SPT) blow count (N), the
cone penetration test (CPT) tip resistance (qc), and small strain Vs. SPT and CPT profiles are
used more often than Vs profiles to evaluate liquefaction triggering (Siegel, 2013).

CRR Relationships

The many relationships available that correlate CRR with measured SPT blow count, CPT
tip resistance, or Vs typically predict the cyclic resistance of the soil for a set of reference
conditions: an initial vertical effective stress of 1 atmosphere, an M 7.5 earthquake, clean sand
(i.e., less than 5% soil grains finer than 0.075 mm), and an area where there is no initial static
shear stress on a horizontal surface (i.e., ground slope is less than 6%). For situations that do not
satisfy these conditions, adjustment factors are required. Differences among the developed CRR
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 87

relationships are due predominantly to the data sets from which they are derived, the adjustment
factors used, and the regression procedure used to develop the CRR curve.
The most commonly used relationships to predict CRR from SPT blow count are those
proposed by Youd and colleagues (2001), Cetin and colleagues (2004), Idriss and Boulanger
(2008), and Boulanger and Idriss (2014). The Youd and colleagues (2001) relationship was
agreed to by consensus at NCEER/NSF workshops in 1996 and 1998, and the others were
published since that time. These CRR relationships were developed from different databases of
case histories, although many of the case histories are common to each database. For each of
these data sets the percentage of cases where liquefaction occurred is between 50% and 55%,
which represents a relatively balanced distribution between “yes” (where liquefaction is believed
to have occurred) and “no” (where liquefaction is believed not to have occurred) cases. The data
for the liquefied and non-liquefied case histories used to develop the Youd and colleagues
(2001), Cetin and colleagues (2004), and Idriss and Boulanger (2008) relationships are shown in
Figure 4.3 along with each proposed CRR relationship. Figure 4.4a shows these three commonly
used CRR relationships on the same graph. The relationships developed by Youd and colleagues
(2001) and Idriss and Boulanger (2008) are similar, while the relationship developed by Cetin
and colleagues (2004) is located below the others, which indicates that for a given SPT blow
count, the Cetin and colleagues (2004) relationship will predict a smaller CRR. Directly
comparing the CRR relationships is misleading, however, because they incorporate different
adjustment factors and use different 𝑟𝑑 relationships. Comparison of the relationships using their
recommended adjustment factors is considered in Box 4.1.
The most commonly used relationships to predict CRR from the CPT tip resistance (qc) are
those developed by Robertson and Wride (1998), Moss and colleagues (2006), and Idriss and
Boulanger (2008). Robertson and Wride (1998) is a slightly modified version of the relationship
recommended by NCEER/NSF workshops in 1996 and 1998 and reported in Youd and
colleagues (2001). The Robertson and Wride (1998) relationship is based on the CRR curve from
Robertson and Campanella (1985), and this curve was derived from the Seed and colleagues
(1985) CRR-SPT relationship by converting SPT blow count into an equivalent CPT tip
resistance. This curve subsequently was compared with CPT data from liquefaction and no
liquefaction case histories and shown to be consistent with the field data. The Moss and
colleagues (2006) and Idriss and Boulanger (2008) relationships were developed directly from
databases of about 200 case histories with measured CPT data. The Moss and colleagues (2006)
relationship is a probabilistic relationship; for deterministic purposes, they recommend use of the
15% probability curve (PL = 15%). Compared with the liquefaction data sets for SPT, the data
sets for the CPT are more heavily weighted toward “yes” cases, with about 70% to 75% of the
cases associated with liquefaction (see Chapter 3). Figure 4.4b plots the predicted CRR as a
function of CPT tip resistance from Robertson and Wride (1998), Idriss and Boulanger (2008),
and Moss and colleagues (2006) for an M 7.5 earthquake and a vertical effective stress of 1
atmosphere. The relationships are similar for CPT tip resistances less than about 50 atmospheres,
but they deviate more substantially at larger values, with the Idriss and Boulanger (2008)
relationship predicting smaller CRR at CPT tip resistances greater than about 100 atmospheres.
Once again, however, this direct comparison may be misleading due to differences in the
adjustment factors used in the different relationships. (See Box 4.1 for a comparison of the
models using their recommended adjustment factors.)
88 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

BOX 4.1
Comparison of Stress-Based Liquefaction Procedures

There are variants of the stress-based method to establish the factor of safety (FS) against
liquefaction. Each variant establishes a relationship between a normalized in situ parameter (e.g., SPT
blow count, CPT tip resistance, or Vs) and the CRR for a standard reference condition (e.g., M = 7.5, v =
1 atmosphere, and clean sand with fines content  5%) and provides recommended adjustment factors to
account for non-reference conditions. Differences among the CRR relationships for the standard
reference condition and differences among the adjustment factors associated with each relationship have
been a source of controversy in geotechnical practice.
Figure 1 compares the computed factors of safety (FSs) versus depth from the different CRR
relationships for a site having a fines content (FC) = 5% and with N 60 = 10, qc = 5 MPa, and Vs = 150 m/s
subjected to an earthquake of M = 7.5 with PGA = 0.1 g, down to a depth of 20 meters. These
comparisons demonstrate the combined effects of the depth-dependent adjustment factors, 𝑟𝑑 (shear
stress reduction factor) and K (the effective overburden stress factor). For the SPT and CPT
relationships, the differences between the FSs computed by the various relationships are not as large as
the differences in the CRR relationships illustrated in Figure 4.4. The differences are more significant for
the Vs relationships, but these relationships were developed from data sets with minimal overlap such that
more variation is expected.
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 89

FIGURE 1 Predicted factor of safety (FS) versus depth for different CRR relationships given M = 7.5 and PGA = 0.1
g. (Yea01: Youd et al. [2001]; Cea04: Cetin et al. [2004]; IB08: Idriss and Boulanger [2008]; RW98: Robertson and
Wride [1998]; Mea04: Moss et al. [2006]; AS00: Andrus and Stokoe [2000]; Kea13: Kayen et al. [2013].) COURTESY:
E. Rathje.
90 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

(a)

(c)
FIGURE 4.3 CRR-SPT relationships developed by (a) Youd and colleagues (2001), (b) Cetin and colleagues (2004),
and (c) Boulanger and colleagues (2012). NOTE: Plotted points represent the earthquake-induced CSR for each
case history corrected to an M 7.5 earthquake, and the initial vertical effective stress, vo, is 1 atmosphere. Lines are
the author-defined boundary curves that separate liquefaction and no liquefaction, and these lines represent the CRR
for M = 7.5 and v = 1 atmosphere as a function of SPT blow count. SOURCE: (a) Youd, T.L., et al. 2001.
Liquefaction resistance of soils: Summary report from the 1996 NCEER and 1998 NCEER/NSF Workshops on
Evaluation of Liquefaction Resistance of Soils. Journal of Geotechnical and Geoenvironmental Engineering 127:817–
833. With permission from ASCE. (b) Cetin, K.O.,et al.. 2004. Standard penetration test-based probabilistic
and deterministic assessment of seismic soil liquefaction potential. Journal of Geotechnical and
Geoenvironmental Engineering 130(12):1314–1340. With permission from ASCE . (c) Boulanger, R.W., D.W.
Wilson, and I.M. Idriss. 2012. Examination and reevaluation of SPT-based liquefaction triggering case histories.
Journal of Geotechnical and Geoenvironmental Engineering 138(8):898–909. With permission from ASCE. These
materials may be downloaded for person use only. Any other use requires prior permission of the American Society
of Civil Engineers. This material may be found at: (a) [http://
ascelibrary.org/doi/abs/10.1061/%28ASCE%291090-0241%282001%29127%3A10%28817%29]; (b) [http://
ascelibrary.org/doi/abs/10.1061/%28ASCE%291090-0241%282004%29130%3A12%281314%29]; (c) [http://
ascelibrary.org/doi/abs/10.1061/%28ASCE%29GT.1943-5606.0000668].
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 91

(a) (b)

(c)

FIGURE 4.4 Comparison of selected CRR predictive relationships: (a) SPT blow count relationships, (b) CPT tip
resistance relationships, and (c) shear wave velocity relationships. COURTESY: E. Rathje.

A commonly used relationship to predict CRR from Vs is the relationship developed by


Andrus and Stokoe (2000), which was recommended for use by the NCEER/NSF workshops
reported in Youd and colleagues (2001). The Andrus and Stokoe (2000) relationship was
developed from a database of 225 case histories, of which about 55% represented liquefaction
cases. More recently, Kayen and colleagues (2013) developed a CRR-Vs relationship based on a
database of 422 case histories; this data set is more heavily weighted to liquefaction cases (about
70%). The Kayen and colleagues (2013) relationship is a probabilistic relationship; for
deterministic purposes, they recommend the use of the 15% probability curve (PL = 15%). Figure
92 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

4.3c plots the predicted CRR for an M 7.5 earthquake and a vertical effective stress of
1 atmosphere as a function of Vs from the Andrus and Stokoe (2000) and Kayen
and colleagues (2013) relationships. The Kayen and colleagues (2013) relationship predicts
larger CRR than Andrus and Stokoe (2000) at shear wave velocities less than about 180
m/s and predicts smaller CRR at larger values of Vs. But again, this comparison may be
misleading due to the different adjustment factors used. (See Box 4.1.)

Adjustment Factors Applied to CRR

Adjustments to the measured in situ resistance parameter and CRR are required
for situations that do not satisfy the standard reference conditions. The situations typically
adjusted in simplified stress-based liquefaction triggering assessment are summarized in Table
4.1. The fines content adjustment is applied to the measured in situ parameter to modify
it to an equivalent clean sand value. Effective stress adjustments are applied to both the in situ
parameter and the CRR. The stress adjustment to the in situ parameter attempts to remove the
effects of vertical effective stress so that the adjusted in situ parameter predominantly reflects
the relative density of the soil. The effective overburden stress factor (K) modifies the CRR to
account for the suppression of dilatancy with increasing effective stress. The shear stress
adjustment (K) is a factor that modifies the CRR to account for the effects of a static shear
stress on the horizontal plane (i.e., non-level ground conditions) on the liquefaction resistance
of the soil.
Each developed CRR relationship incorporates its own set of adjustment factors. The
equations for these adjustment factors have become increasingly complex as a
consequence of efforts to incorporate the effects of different potential conditions that may
influence them. This extensive fine-tuning of the adjustment factors—in particular, for
conditions not constrained by field case history data—has complicated the
“simplified” method of liquefaction evaluation. Adjustment factors need to be carefully
evaluated as additional case history data become available, and modifications to the
adjustment factors that increase their complexity should result in decreasing uncertainty.
There are two basic approaches to developing adjustment factors: one, they can
be developed using theoretical constraints from soil mechanics principles and experimental data,
or two, they can be derived directly in the regression for the CRR relationship.
The liquefaction case history databases generally do not include enough data
either at effective stresses corresponding to depths less than 1.5 meters or in excess of 15
meters or at small (e.g., M < 6.0) and very large (e.g., M > 8.0) earthquake magnitudes
to constrain the adjustment factors adequately within a regression analysis. Additionally,
using regression analysis alone may result in unintended trends in the relationship: for
example, the MSF within the Moss and colleagues (2006) regression equation results
in a weak magnitude dependence for CRR. Using soil mechanics principles and
experimental data to constrain adjustment factors beyond the range over which they are
constrained by case history data is a fundamentally more sound approach than is simple
extrapolation of regression equations. Documenting field case histories that can constrain
adjustment factors outside the current bounds is a priority.
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 93

TABLE 4.1 Situations Typically Requiring Adjustment Factors in CRR Calculation


Applied to In Situ
Adjustment Parameter? Applied to CRR?
a a
Fines Content Yes No

Effective Stress Yes Yes

Static Shear Stress No Yes


a
The adjustment for fines content may be applied to either the in situ parameter (i.e., SPT N value, CPT qc, shear
wave velocity) or the CRR.

Adjustment for Fines Content

The fines content influences both the measured in situ parameter and the CRR, but it is
difficult to separate the two effects. Early CRR relationships (see, e.g., Seed et al., 1985)
included curves for different fines contents. Most current CRR relationships only provide a clean
sand curve to be used with an equivalent clean sand in situ parameter: for example, (N1)60cs,
qc1Ncs. The shear wave velocity CRR relationships do not show a strong influence of fines
content. The developers of each CRR relationship recommend their own method to adjust for
fines content. Most fines content adjustments have been developed directly from the case history
databases, and they use the measured fines content from soil samples as input. Traditional CPT
soundings and Vs measurements, however, do not provide soil samples for the direct
measurement of fines content. The adjustment factors for CPT use measurements of tip
resistance and skin friction directly to adjust the CPT tip resistance (Robertson and Wride 1998;
Moss et al., 2006), often using the soil behavior type index (Robertson, 1990).
The different approaches to correct for fines content could result in discrepancies among
different liquefaction triggering relationships. Research is needed to better evaluate these
different approaches. The consistency between the methods that use direct measurement of fines
content and those that infer fines content from CPT measurements needs to be investigated. Until
the indirect methods are fully vetted, it is good practice to measure fines content at a site directly
and to develop site-specific relationships between fines content and the soil behavior type index.

Adjustments for Effective Stress

The stress adjustment factors applied to correct in situ penetration test parameters to a
vertical effective stress of 1 atmosphere were developed from data from large-scale calibration
chamber tests (see, e.g., Marcuson and Bieganousky, 1977), field tests (see, e.g., Skempton,
1986), and theoretical solutions (see, e.g., Salgado et al., 1997). In the stress ranges where little
to no experimental or field data exist, the relationships for the adjustment factors were either
extrapolated from stress ranges where more data existed and/or extrapolated based on soil
mechanics principles. The relationships for the adjustment factors associated with the different
variants of the simplified procedure can have a considerable influence on the predicted
liquefaction potential, particularly at higher effective overburden stresses (e.g., depths greater
than approximately 20 meters).
94 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

The stress adjustment factor applied to the CRR (i.e., K, defined as the ratio of the CRR for
a given vertical effective stress to the CRR for a vertical effective stress of 1 atmosphere for the
same soil) was first proposed by Seed (1983). Values of K have been derived from either
laboratory data and soil mechanics principles (see, e.g., Hynes and Olsen, 1999; Youd et al.,
2001; Idriss and Boulanger, 2008, 2010; Dobry and Abdoun, 2015) or directly from regression of
the liquefaction database (see, e.g., Cetin et al., 2004; Moss et al., 2006; Kayen et al., 2013). It is
important to note that none of the liquefaction case histories include critical layers in which the
vertical effective stress was greater than 2 atmospheres, which corresponds to a depth below
ground level of approximately 20 meters. Most cases correspond to vertical effective stresses less
than 1 atmosphere (see Chapter 3, Table 3.2). The values of K developed by different
researchers generally agree over effective stresses ranging 0.5 to 1.2 atmospheres—the range
well represented in the case history database. Recommended values for Kσ begin to deviate
beyond that range.
The depth of the water table below the ground surface can be important
in liquefaction potential assessments. Fluctuations in the water table depth may occur
on timescales of years (e.g., in areas of groundwater injection or withdrawal or during periods
of drought), seasonally, or daily (e.g., in a port and harbor environment). The water table
depth may influence both the extent of the saturated zone that is susceptible to liquefaction
and the value of the effective stress at any particular point below its surface.
Observations made following the 2011 earthquake in Christchurch, New Zealand, suggest
unsaturated zones of sandy soil are likely below the water table where short-term (e.g.,
daily) fluctuations in the water table occur (Stokoe et al., 2014a). In such zones, the potential
for liquefaction triggering is mitigated, but this may not be true in high-permeability soils such as
gravel and rockfill.
The change in effective stress due to water table fluctuations can impact the
cyclic resistance of the soil, particularly at shallow depths where small changes in the water
table can result in relatively large changes in the effective stress used to calculate the
cyclic stress ratio. For example, if the water table changes from four to two meters
beneath the ground surface, the effective stress at a depth of five meters can decrease from
approximately 88 kPa to 64 kPa, a decrease of more than 25%. At present, there is no
consensus and little guidance on what values for water table level to use when the water
table is known to fluctuate. The absence of such guidance may lead the geotechnical
engineer to use historic high water tables (if such data are available) or the mean high
water level (e.g., for port and harbor projects, because the higher the water table the greater
the potential for triggering liquefaction). In a true probabilistic assessment of liquefaction
potential, the depth of the water table could be treated as a random variable and assigned a
mean and standard deviation to capture the fluctuations. Results from such analyses could be
used to establish guidance on how to appropriately account for the water table level in a
deterministic analysis. In the absence of such guidance, professional judgment is necessary.
The K relationships deviate at small effective stresses because there is a debate about
whether it is advisable to cap the maximum value of K and, if so, at what value to cap it. At
effective stresses significantly larger than 1 atmosphere, recommended values of Kσ represent
significant extrapolations from the liquefaction case history databases, and the values
recommended by different researchers deviate significantly from each other. Given the
uncertainty involved in any extrapolation, it is important that experimental data, soil mechanics
principles, or both be used to constrain K outside of the range represented in the database. As an
example of the problems associated with using only field case histories to constrain Kσ within the
regression, some CRR relationships show almost no stress dependency because the case history
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 95

data they employ do not extend across a large enough stress range to constrain the
influence of effective stress on CRR: this was shown by Moss and colleagues (2006)
and Kayen and colleagues (2013). Nevertheless, experimental data and soil mechanics
principles do indicate that effective stress influences CRR. Consequently, Kayen and colleagues
(2013) do recommend that stress adjustment factors from other studies be applied to the
computed CRR from their relationships.
One possible reason why the existing liquefaction databases do not include enough data at
large effective stresses to constrain the stress corrections is that liquefaction at depth may not
produce a surface manifestation of liquefaction, and surface manifestation of liquefaction is the
primary means of identifying sites where liquefaction has occurred. In engineering practice,
however, liquefaction triggering analysis is performed routinely for soil at significant depth
under large effective stress. Although the measured in situ test parameters for these deep soils
may be far in excess of any uncorrected values in the liquefaction databases, a liquefaction
triggering analysis may show such soils to be highly liquefiable after application of the various
stress adjustment factors. Data from laboratory experiments performed at large effective stresses
have been used by some investigators to constrain the stress adjustment factors (see, e.g., Seed
and Harder, 1990) at vertical effective stresses larger than 1 atmosphere, and some investigators
have also incorporated soil mechanics principles (e.g., principles from critical state soil
mechanics) to constrain these adjustment factors (see, e.g., Boulanger, 2003b). More recently,
Montgomery and colleagues (2012) compiled an expanded database of experimental results at
large effective stress and evaluated the Kσ relationships of Youd and colleagues (2001) and Idriss
and Boulanger (2008). They concluded that the Youd and colleagues (2001) relationship
predicted values of Kσ that were 10% to 15% smaller than the data and that the Idriss and
Boulanger (2008) relationship predicted values of Kσ that were 0% to 10% larger. A thorough
examination of all of the available experimental data from laboratory and centrifuge tests
regarding stress adjustment factors at both small and large effective stresses is warranted in an
effort to provide more insight about the values of K at large effective stresses.

Adjustment for Initial Static Shear Stress

The initial static shear stress on the horizontal plane present in the soil before earthquake
shaking also influences the CRR and is taken into account through the multiplicative shear stress
adjustment factor, K, introduced by Seed (1983). A parameter  is defined as the ratio of the
initial static shear stress to the normal effective stress on the horizontal plane, and K represents
the ratio of the CRR for a given value of  to the CRR for  = 0. As the CRR relationships
represent level or gently sloping ground (e.g., ground slope less than 6%), the CRR obtained
from a CRR relationship represents  = 0 (i.e., level ground) conditions and must be adjusted if
the in situ  is larger than zero (e.g., slopes and in and under earth dams).
Insights into the effects of an initial static shear stress on the CRR have been obtained
predominantly from laboratory tests. These experimental data show that K is influenced
not only by  but also by the state of the soil as expressed both by its relative density and
the effective confining pressure and by the criterion used to define liquefaction in the
experiments. Importantly, the presence of an initial static shear stress appears to increase the
CRR for denser soils and to decrease it for looser soils. Seed and Harder (1990) compiled
a wide range of laboratory test data from the 1970s and 1980s and used a strain-
based definition of liquefaction
96 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

to develop recommended values for K that vary with  and relative density. Harder and
Boulanger (1997) revised those recommended values in light of new experimental data
(Boulanger et al., 1991; Boulanger and Seed, 1995), including data on the effect
of multidirectional shaking and the use of a smaller strain criterion to define liquefaction
triggering. Nevertheless, the consensus from the NCEER/NSF liquefaction workshops
(Youd et al., 2001) was not to recommend values of K because of the large scatter in the
experimental data and to recommend that additional research on this subject was needed.
Since the NCEER/NSF workshops, the theoretical framework of critical state soil mechanics
has been applied to the experimental data to explain the combined effects of relative density and
effective stress on the K relationships (Boulanger, 2003a). Additionally, new experimental data
have been generated (see, e.g., Kammerer, 2002; Wu et al., 2003; Cetin and Bilge, 2014), much
of it focused on multidirectional shaking using a wide variety of stress paths. But there is still no
consensus on recommended K factors for use in engineering practice. For projects that
require the effects of an initial static shear stress to be considered (e.g., large dams), the
most commonly used K factors are those from Boulanger (2003a). These empirical factors
were developed with guidance from soil mechanics principles and were compared with a
small set of experimental data for three different sands. A favorable comparison was
observed, but only a limited amount of data was considered. Therefore, additional research
is needed for clear resolution of this topic. This research needs to take advantage of centrifuge
testing, for example, as well as physics-based numerical modeling (e.g., discrete element
analysis) to improve understanding of this effect.

Use of Consistent Adjustment Factors in Engineering Practice

It is critical to note that the liquefaction triggering adjustment factors, including the
magnitude scaling factor (MSF) and 𝑟𝑑 factors discussed earlier, are used in the development of
the CRR triggering relationships. These factors are used to adjust the CSR (i.e., demand) for each
case history to the reference conditions (i.e., v = 1 atmosphere, M = 7.5, clean sand
conditions) before plotting it on graphs such as those shown in Figure 4.3. Therefore, the
adjustment factors directly influence the location of the derived CRR relationship. To use a CRR
relationship in engineering practice appropriately, adjustment factors consistent with those used
in the development of the CRR relationship must be employed. Engineers should not mix and
match adjustment factors from different CRR relationships. This was not commented on during
previous efforts to develop consensus (i.e., Youd et al., 2001). Thus, the use of site response
analysis to estimate the earthquake-induced CSR is appropriate if, and only if, the 𝑟𝑑 factors used
in the development of the CRR curve are unbiased relative to the values expected at liquefaction
sites. The 𝑟𝑑 factors developed by Cetin and Seed (2004) are considered by the present study
committee to be unbiased because they analyzed about 50 well-characterized liquefaction sites
and used a large number of earthquake motions. Therefore, site response analysis may be
employed for the computation of the earthquake-induced CSR in conjunction with the CRR
relationships developed using the Cetin and Seed (2004) 𝑟𝑑 factors (i.e., Cetin et al., 2004; Moss
et al., 2004; and Kayen et al., 2013).
To avoid compounding uncertainties in CRR relationships developed in the future, unbiased
estimates of all parameters need to be used. This includes using site response analysis to estimate
the earthquake-induced CSR for all of the liquefaction case histories employed in development
of a new CRR relationship.
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 97

Comparisons Between CRR Relationships

There are controversies related to the differences among various CRR relationships. Idriss
and Boulanger (2012) state that certain misinterpretations and inconsistencies in the Cetin and
colleagues (2004) database are responsible for the differences between the CRR relationships for
the SPT shown in Figure 4.4a. Preliminary results provided to the committee by Cetin (written
communication, June 29, 2014),3 however, indicate that the inconsistencies identified by Idriss
and Boulanger (2012) do not noticeably alter the location of the Cetin and colleagues (2004)
CRR curve. The information provided to the committee by Cetin suggests that the differences in
the Cetin and colleagues (2004) and Idriss and Boulanger (2008) CRR relationships are due
primarily to differences in 𝑟𝑑 and Kσ. The results shared with the committee by Cetin need to be
evaluated carefully by the technical community once published in the peer-reviewed literature.
Additionally, when using adjustment factors consistent with the development of each
relationship under conditions where both relationships are well constrained by data (e.g., depths
less than 20 meters), the differences between the Cetin and colleagues (2004) and Idriss and
Boulanger (2008) triggering approaches are not as significant as suggested by the CRR
relationships alone, as illustrated in Box 4.1. At depths greater than 20 meters, however, the FS
evaluated using these methods diverges (Griffiths and Cox, 2012).

Comparison of Triggering Evaluation Methods

Several liquefaction triggering evaluation procedures are commonly used in practice (see,
e.g., Robertson and Wride, 1998; Andrus and Stokoe, 2000; Youd et al., 2001; Cetin et al., 2004;
Moss et al. 2006; Idriss and Boulanger, 2008; Kayen et al., 2013; Boulanger and Idriss, 2014).
As discussed throughout this chapter, each procedure has advantages and disadvantages. For
example, a notable strength of the approach developed by Idriss and Boulanger (2008) is the use
of laboratory and soil mechanics principles to extrapolate the applicable range of the procedure
beyond that constrained by case study data. And, a notable strength of the Cetin and colleagues
(2004) procedure is the use of the unbiased estimator for 𝑟𝑑 (i.e., for adjusting the earthquake-
induced shear stress computed using equation 4.1 to account for the nonrigid response of the soil
column). Differences between various methods that are soundly based may be considered to be
epistemic uncertainty. As such, the engineer may apply various methods and weight the results,
wherein the weighting factors are developed based on professional judgment on a case-by-case
basis depending on the specifics of the problem. The sum of all weights should be 1.0.

Probabilistic CRR Relationships

Various CRR relationships have been developed within a probabilistic framework. Most of
these use well-known statistical methods like logistic regression or discriminant analysis (see,
e.g., Liao et al., 1988; Youd and Noble, 1997a) to distinguish between liquefaction and no-
liquefaction categories. It is not widely appreciated, however, that these regression techniques
provide the probability of observing the collected liquefaction data set given that liquefaction has
3
Available by request from https://www8.nationalacademies.org/cp/ManageRequest.aspx?key=49573.
98 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

occurred. These techniques do not provide the probability of liquefaction given the observed in
situ parameters at a site. Some researchers have used Bayesian techniques to avoid this problem
(see, e.g., Cetin et al., 2004; Moss et al., 2006), usually assuming a non-informative prior
probability that indicates liquefaction and non-liquefaction are equally probable at the site being
evaluated, or employing non-informative prior probabilities in developing the model coefficients
themselves. This assumption may not be reasonable for some sites because information is
available that makes liquefaction more or less likely (e.g., geologic age, depositional
environment, evidence of prior liquefaction), and this prior probability information could be used
to enhance a liquefaction analysis. Plotting the CRR curves associated with the likelihood ratio
rather than probability would enable the user to deal with prior and posterior probabilities
directly within a Bayesian framework (see Box 4.2).

Effect of Geologic Age

Chapter 3 (“Influence of Soil Age on Liquefaction Potential”) provided a description of the


effects of geologic age on soils. The effects of age often manifest themselves as an increase in
CRR with time under confinement (e.g., Seed, 1979). Ideally, any increase in liquefaction
resistance with age would manifest itself in larger values of the in situ parameters used to
evaluate CRR (e.g., SPT blow count, CPT tip resistance, or Vs). Penetration resistance measured
by the SPT and the CPT, however, may not be very sensitive to the effects of aging (Andrus et
al., 2009) because these tests induce large strains in the soil that may destroy the beneficial
effects of aging. Andrus and colleagues (2009) have suggested that Vs, which reflects the small-
strain stiffness of the soil, is more sensitive to the effects of aging and therefore can be used to
account for aging effects.
Significant uncertainty remains regarding the physical mechanisms that increase the
liquefaction resistance with time, although secondary compression and biogeochemical processes
may be involved. Additionally, the effect of aging on the liquefaction resistance of field deposits
has been studied in only a few geographic areas (e.g., Charleston, South Carolina, and
Christchurch, New Zealand). Further research is needed to understand fully the effects of aging
on liquefaction resistance and to develop rational and defensible methods for taking it into
account.
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 99

BOX 4.2
Probabilistic Liquefaction Triggering Plots

Most statistical procedures for separating data into categories are versions of either discriminant
analysis or logistic regression (Klecka, 1980; Pampel, 2000; Pohari et al., 2004). Such analyses applied
to liquefaction give the probability of observing the data (i.e., in situ parameter and CSR) given that
liquefaction occurs, P[d | L ], where “d” represents the data and “L” indicates that liquefaction occurs. This
is not the same as the probability that liquefaction occurs given the data, P[ L | d ].
Several researchers (see, e.g., Juang et al. 2002; Cetin et al. 2004) incorporate Bayesian updating
into their analyses, which requires a prior estimate of the probability of liquefaction (P 0[ L ]). In the
absence of other information, it is common to assume a “non-informative prior,” as has been done in all of
the published liquefaction analyses that invoke Bayesian updating. It means equal prior probabilities of
liquefaction and non-liquefaction.
Published probabilistic liquefaction charts can be used to develop a posterior estimate of the
probability of liquefaction, P1[ L | d ], accounting for an informed prior if Bayes’ theorem is written as
likelihood ratios (LRs) and odds. For the liquefaction problem, the LR is the probability of observing the
data given that liquefaction occurred, P[ d | L ], divided by the probability of observing the data given that
liquefaction did not occur, P[d|nL]:

𝑃[𝑑|𝐿]
𝐿𝑅 = (1)
𝑃[𝑑|𝑛𝐿]

The odds are the ratio of the probabilities of a positive result (e.g., liquefaction) to a negative result (e.g.,
no liquefaction). The odds of liquefaction are:

𝑃[𝐿]
𝑜𝑑𝑑𝑠[𝐿] = (2)
𝑃[𝑛𝐿]

Bayes’ theorem can be written as (see Baecher and Christian, 2003; Gelman et al., 2014):

𝑃1 [𝐿|𝑑] 𝑃0 [𝐿] 𝑃[𝑑𝐿]


= ∙ (3)
𝑃1 [𝑛𝐿|𝑑] 𝑃0 [𝑛𝐿] 𝑃[𝑑|𝑛𝐿]

Substituting the definitions of the odds and likelihood ratios into (3) results in:

𝑝𝑜𝑠𝑡𝑒𝑟𝑖𝑜𝑟 𝑜𝑑𝑑𝑠[𝐿|𝑑] = 𝑝𝑟𝑖𝑜𝑟 𝑜𝑑𝑑𝑠[𝐿|𝑑] ∙ 𝐿𝑅 (4)

Since P1[ nL | d ] = 1 – P1 [ L | d ], we can relate P1[ L | d ] to the posterior Odds:

𝑝𝑜𝑠𝑡𝑒𝑟𝑖𝑜𝑟 𝑜𝑑𝑑𝑠[𝐿|𝑑]
𝑃1 [𝐿|𝑑] = (5)
1+𝑝𝑜𝑠𝑡𝑒𝑟𝑖𝑜𝑟 𝑜𝑑𝑑𝑠[𝐿|𝑑]

We can now use equations (4) and (5) to develop posterior estimates of the probability of
liquefaction. Assume that the data for a site fall on a line labeled with 15% probability of liquefaction (i.e.,
P1[ L | d ] = 0.15) and that the lines were developed using a non-informative prior. The
posterior probability of non-liquefaction (P1[ nL | d ]) must be 85%, so the posterior odds for this
point must be 0.15/0.85 = 0.1765. The prior odds (P0 [ L] / P0[ nL ]) for a non-informative prior were 1.0,
so LR must be 0.1765. Now, consider a site for which geological and hydrological evidence (e.g., from an
analysis such as described in Zhu et al., 2014) indicate a prior probability of liquefaction of 80%. The prior
odds are now 0.80/0.20 = 4.0. Using equation (4) the posterior odds of liquefaction are therefore
(4.0)(0.1765) = 0.7060, and using equation (5) the updated probability of liquefaction is then
(0.7060)/(1+0.7060) = 0.4139, or 42%. This is a significant reduction from the value of 80%; it is also
much larger than the 15% suggested by the non-informative prior.
This example shows that the current probabilistic liquefaction charts based on Bayesian updating and
a non-informed prior can be used to compute a posterior estimate of the probability of liquefaction
100 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

that accounts for an informed prior. Reporting the likelihood ratio for the CRR curves directly, however,
would make this computation easier and is preferable to relying on curves for constant values of
probability based on non-informative priors.

Fine-Grained Soils

In the past, if more than 50% of a soil’s constituents (by weight) passed the #200 (0.075
mm) sieve (i.e., fine-grained soils), the soils were considered non-liquefiable. It is now
commonly recognized that low plasticity fine-grained soils are susceptible to liquefaction. It is
also recognized that soils with less than 50% passing the #200 sieve may not be susceptible to
liquefaction if the percentage passing through the #200 sieve and the plasticity index (PI) of the
soil that passes through are above certain threshold values. This may be true even if those soils
are classified as a sand, or possibly a gravel, under the Unified Soil Classification System.
Various criteria for liquefaction susceptibility of soils have been proposed involving grain size
distribution, water content, Atterberg limits (e.g., liquid limit [LL] and PI), and CPT indices (see,
e.g., Wang, 1979; Seed and Idriss, 1982; Iai et al., 1986, 1989; Robertson and Wride, 1998;
Andrews and Martin, 2000; Youd et al., 2001; Seed et al., 2003; Bray and Sancio, 2006;
Boulanger and Idriss, 2006).
As discussed in Chapter 3, many practicing engineers still use the “Chinese criteria” when
assessing the potential for liquefaction triggering in soils containing fine-grained (silt and clay-
sized) particles. Seed and Idriss (1982) stated that if clay size soils meeting the criteria (see
Chapter 3) plotted above the A-line on a Casagrande plasticity chart, liquefaction characteristics
should be determined through testing; otherwise, the soils were not considered vulnerable to
liquefaction.
Based on differences in index tests in the United States and China (Koester, 1992), Andrews
and Martin (2000) proposed the modification to the Chinese criteria presented in Table 4.2. Also,
Moss and Chen (2008) state that the 2001 Chinese Building Code (CNS, 2001) used a slight
variation of the original Chinese criteria, specifying the clay fraction limit (i.e., the maximum
percent by weight finer than 0.005 mm) above which the soil is considered not susceptible to
liquefaction as a function of the intensity of ground shaking. The CNS (2001) specifies the clay
fraction as 10%, 13%, and 16% for Chinese Intensity 7, 8, and 9, respectively, and Moss and
Chen (2008) note that Chinese Intensity 7 through 9 is approximately equal to Modified Mercalli
Intensity VI through X.
Although the Chinese criteria are still used, more recent investigators have dismissed them
as insufficient or simply incorrect. Alternative criteria commonly used in practice in the United
States are those developed by Seed and colleagues (2003), Bray and Sancio (2006), Boulanger
and Idriss (2006), and Robertson and Wride (1998). The first three criteria are based on index
properties determined from laboratory tests. Robertson and Wride (1998) define a soil behavior
type index, Ic, in terms of normalized values of CPT tip resistance and friction ratio.
The liquefaction susceptibility criterion proposed by Robertson and Wride (1998) is based
on the empirical classification of soils with Ic ≤ 2.6 as sand and sand mixtures (see Figure 4.5)
and thus are assumed to be susceptible to liquefaction. Robertson and Wride (1998) suggest that
the CPT version of the simplified stress-based liquefaction procedure can be used to evaluate the
liquefaction potential of these soils. Soils with Ic > 2.6 are classified empirically as clayey silt,
silty clay, clay, or organic soils (see Figure 4.5). Robertson and Wride (1998) state that soils with
Ic > 2.6 should be sampled and evaluated for liquefaction susceptibility based on laboratory
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 101

measured index properties and cyclic shear testing. Youd and colleagues (2001) lowered the
threshold for sampling and testing to evaluate liquefaction potential to Ic > 2.4. The original
Robertson and Wride (1998) criterion (i.e., soils with Ic ≤ 2.6 are assumed to be liquefiable and
can be evaluated using the simplified procedure) was adopted by the Earthquake Commission
(EQC) in New Zealand as the criterion for determining susceptibility to liquefaction triggering
for sites in Christchurch (Tonkin and Taylor, 2013).

TABLE 4.2 Modified Chinese Criteria Proposed by Andrews and Martin (2000)
Liquid Limit < 32 (1) Liquid Limit > 32
Clay Content < 10% (2) Susceptible Further Studies Required
(Considering plastic non-
clay-sized grains such as
mica)

Clay Content > 10% Further Studies Required Not Susceptible


(Considering non-plastic clay-
sized grains such as mine and
quarry tailings)

(1) Liquid Limit determined by Casagrande-type percussion apparatus.


(2) Clay defined byas grains finer than 0.002 mm.

FIGURE 4.5 Correlation relating soil behavior type index (Ic) and Apparent Fines Content. SOURCE: Robertson, P.K.,
and C.E. Wride. 1998. Evaluating cyclic liquefaction potential using the cone penetration test. Canadian Geotechnical
Journal 35(3):442–459. © 2008 Canadian Science Publishing, with permission.
102 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Based on a combination of post-earthquake observational data and laboratory tests, Seed


and colleagues (2003), Bray and Sancio (2006), and Boulanger and Idriss (2006) proposed
liquefaction susceptibility criteria and liquefaction potential assessment procedures based upon
the percent fines and Atterberg limits of the soil. The case history data supporting the
development of these criteria are discussed in Chapter 3 in the section on liquefaction of fine-
grained soils.
Seed and colleagues (2003) proposed the screening criteria shown in Figure 4.6 which
defines three zones of susceptibility to what they describe as “classic cyclic liquefaction”—the
loss of strength and stiffness due to cyclic pore-pressure generation. These zones include soils
that are potentially liquefiable (Zone A in Figure 4.6); soils that may be potentially liquefiable
(Zone B of the figure); and soils not susceptible to “classic cyclic liquefaction.” Seed and
colleagues (2003) explain that soils classifying as not susceptible to “classic cyclic liquefaction”
could exhibit what they called “sensitivity” to liquefaction: strength loss resulting from
monotonic shearing or remolding due to large, monotonic (unidirectional) shear displacements.
PI is used in place of percent clay fines in the Seed and colleagues (2003) criteria, as is done in
the Chinese criteria. Water content (wc) and LL are still part of the Seed and colleagues (2003)
criteria. Those criteria apply to fine-grained soils with either a fines content (FC) equal to or
greater than 20% if the PI is equal to or greater than 12, or an FC equal to or greater than 35%
and a PI less than or equal to 12. Soils considered potentially susceptible to “classic cyclic
liquefaction” (Zone A in Figure 4.6) can be evaluated using the simplified procedure. Soils
falling in Zone B and thus possibly susceptible to “classic cyclic liquefaction” need verification
of their liquefaction susceptibility—beyond evaluation through the simplified procedure—
through sampling and laboratory testing. Zone C soils, although generally not considered to be
susceptible to “classic cyclic liquefaction,” may be sensitive to liquefaction.

FIGURE 4.6 Liquefaction susceptibility criteria proposed by Seed and colleagues (2003) showing zones of
susceptibility to what they described as “classic cyclic liquefaction.” SOURCE: Nisee/Peer, University of California,
Berkeley.
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 103

FIGURE 4.7 Liquefaction susceptibility criteria proposed by Bray and Sancio (2006) showing zones of susceptibility to
liquefaction triggering. SOURCE: Bray, J.D., and R.B. Sancio. 2006. Assessment of the liquefaction susceptibility of
fine grained soils. Journal of Geotechnical and Geoenvironmental Engineering 132(9):1165–1177. With permission
from ASCE. This material may be downloaded for personal use only. Any other use requires prior permission of the
American Society of Civil Engineers. This material may be found at
[http://ascelibrary.org/doi/abs/10.1061/%28ASCE%291090-0241%282006%29132%3A9%281165%29].

Bray and Sancio (2006) also identified three zones of susceptibility for soils containing fine-
grained particles, shown graphically in Figure 4.7: susceptible, moderately susceptible, and not
susceptible. Soils that fall in their “not susceptible” zone are not considered to be susceptible to
liquefaction triggering but may be sensitive. According to Bray and Sancio (2006), the cyclic
behavior of fine-grained soils is controlled by numerous factors. They recommend that soils that
soils categorized as “susceptible” and “moderately susceptible” be sampled and tested to
ascertain their susceptibility to liquefaction and their strain potential. This recommendation
assumes that soils in the susceptible zone have enough fines such that they can be sampled and
tested without excessive disturbance, which is not unreasonable given that the data on which
Bray and Sancio (2006) based their criteria included mostly soils with an FC ≥ 35%. Their
recommendation to sample and test all soils with FC ≥ 35% differs from the finding of Seed and
colleagues (2003)—that the liquefaction potential of soils in Zone A may be evaluated using the
simplified procedure.
The criterion proposed by Boulanger and Idriss (2006) is solely a function of the PI. As
discussed in Chapter 3, they use their criteria to distinguish between soils that are susceptible to
liquefaction triggering (i.e., they exhibit “sand-like” behavior) and those that are susceptible to
some strength loss due to cyclic loading but are not susceptible to the loss of stiffness and
strength and cyclic pore-pressure generation characteristic of what Seed and colleagues (2003)
referred to as “classic cyclic liquefaction” (i.e., they exhibit “clay-like” behavior). Boulanger and
Idriss (2006) state that soil behavior transitions from sand-like to clay-like over the PI range of
approximately 3 to 8, but they recommend PI = 7 as a conservative boundary between the two
behaviors (see Figure 4.8). The distinction between the two types of behavior defined by
Boulanger and Idriss (2006) relates not only to differences in the cyclic response of these soils
but also to differences in the appropriate testing and analysis procedures used to evaluate the
dynamic response characteristics of these soils (Armstrong and Malvick, 2014, 2015). The
simplified stress-based liquefaction evaluation procedure is appropriate only for soils having PI <
7 (i.e., sand-like behavior). On the contrary, soils having PI ≥ 7 (i.e., clay-like behavior) may be
susceptible to cyclic softening or may be sensitive and should be evaluated accordingly, but the
104 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

simplified stress-based liquefaction evaluation procedure should not be used to evaluate the
dynamic response characteristics of these soils.

FIGURE 4.8 Liquefaction susceptibility criteria proposed by Boulanger and Idriss (2006) based on plasticity
indices.SOURCE: Boulanger, R.W., and I.M. Idriss. 2006. Liquefaction susceptibility criteria for silts and
clays. Journal of Geotechnical and Geoenvironmental Engineering 132(11):1413–1426. With permission
from ASCE. This material may be downloaded for personal use only. Any other use requires prior permission of the
American Society of Civil Engineers. This material may be found at
[http://ascelibrary.org/doi/abs/10.1061/%28ASCE%291090-0241%282006%29132%3A11%281413%29].

When comparing the Seed and colleagues (2003), the Bray and Sancio (2006), and the
Boulanger and Idriss (2006) criteria, note that Seed and colleagues (2003) use a combination of
FC and PI to define the boundary between susceptible and not susceptible soils (i.e., fines-
controlled soils are those having FC > 20% and PI > 12 or FC > 35% and PI < 12). In contrast,
Bray and Sancio (2006) apply their criteria to all soils having FC > 35%, and Boulanger and
Idriss (2006) state that their criterion can be used for soils having FC less than 50% if it can be
shown that the soil behavior is controlled by fine-grained fraction. These three sets of criteria
also differ in the required input properties and their intended use. The criteria proposed by Seed
and colleagues (2003) and Bray and Sancio (2006) require PI, LL, and water content (wc), while
the criterion proposed by Boulanger and Idriss (2006) requires only PI. Laboratory testing of
undisturbed samples in all three criteria for a class of soils that fall between susceptible and not
susceptible, though the criteria for these intermediate soil classes vary among the three sets of
criteria. Laboratory testing to evaluate the liquefaction triggering of undisturbed soil specimens
is discussed in Chapter 5.
Seed and colleagues (2003) and Boulanger and Idriss (2006) state that the simplified method
(with appropriate fines content corrections) can be used for the soils they identify as susceptible
to liquefaction. Bray and Sancio (2006) make no statements about the applicability of the
simplified stress-based liquefaction evaluation procedure for soils that fall into their susceptible
(to liquefaction) zone. For establishing whether or not laboratory testing is needed to assess if a
soil is liquefiable, however, their recommendations apply only to soils with FC greater than 35%,
suggesting that the simplified method not be used and that laboratory testing is required when the
FC is greater than 35%.
While similar conclusions regarding liquefaction susceptibility may be reached from these
three sets of criteria for most soils, the varying criteria inevitably will yield different conclusions
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 105

about the liquefaction susceptibility for some fine-grained soils. This potential for conflicting
conclusions is exacerbated by the relatively poor accuracy and precision associated with
measurements of Atterberg limits (e.g., LL and PI) in low-plasticity soils. As a result, all criteria
should be applied with caution in any liquefaction assessment. Furthermore, additional field data
on fine-grained soils, including soils with an FC between 35% and 50%, are necessary to resolve
the conflicts between these sets of recommendations. As noted in Chapter 3, data need to include
those for fine-grained. Additionally, because these soils generally are amenable to undisturbed
sampling, field data need to be supplemented with laboratory testing of intact samples to refine
and/or validate the criteria limits.

Gravelly Soils

Loose gravelly soils have liquefied in earthquakes (see, e.g., Youd et al., 1985). Gravelly
soils may consist of gravel-sized particles floating in a matrix of finer particles (e.g., sands, silts,
and clays) or may consist of a gravel matrix in which the gravel particles are in contact and finer
particles may fill in part of the pore space. Gravel soils that do not contain a significant finer-
grained component are believed to be less susceptible to liquefaction due to their large hydraulic
conductivity, which allows porewater pressures to dissipate quickly during earthquake shaking
(Kramer, 1996). Nevertheless, layers of gravel with few fines confined between low-
permeability layers can liquefy in an earthquake because the low-permeability layers impede
drainage (Stokoe et al., 1988).
It is difficult to characterize the liquefaction resistance of gravelly soils using the SPT or the
CPT because the large size of the particles relative to the SPT sampler or the CPT cone tip can
increase the measured penetration resistance in a nonrepresentative manner. Different types of in
situ tests, such as the Becker penetration test (BPT; Harder and Seed, 1986), various large
penetration tests (LPTs; Daniel et al., 2003), and large dynamic penetration tests (DPTs; Cao et
al., 2013), have been developed to circumvent these problems. To minimize the effects of large
particles, the BPT uses a diesel hammer to drive a 17-cm diameter closed-end pipe, while LPTs
drive large samplers (7 to 14 cm diameter) using larger hammers than the SPT hammer. A DPT
developed in China (Cao et al., 2013) drives a 7.4-cm diameter cone tip with a large hammer and
reports the number of blows to advance 10 cm. Both the BPT and the LPTs measure the number
of blows to advance 30 cm and attempt to convert that number to an equivalent SPT blow count
(see, e.g., Harder and Seed, 1986).
Research has shown that the BPT is influenced by variable hammer energy and efficiency
(Sy and Campanella, 1994) as well as shaft friction (Ghafghazi et al., 2014). Details of the test
setup influence the LPT (Daniel et al., 2003). As a result, there is significant scatter in the
conversions of BPT and LPT penetration resistances to the SPT (see, e.g., Youd et al., 2001),
which increases the uncertainty in their use in liquefaction evaluations. An instrumented BPT is
being developed (DeJong et al., 2014); Ghafghazi et al., 2014) that measures the energy at the tip
of the BPT penetrometer, which can be used to improve the BPT-SPT conversion.
Further improvements to the LPT and BPT test procedures and their conversion to
equivalent SPT blow counts are still needed for these methods to be considered reliable means of
assessing liquefaction resistance. Direct evaluation of the liquefaction resistance of gravels using
these test methods is preferable to converting the measurements to SPT blow count, but the case
history data required to do this are not available at the present time. This direct approach has
106 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

been applied on a limited scale to the Chinese DPT (Cao et al., 2013). Consideration should also
be given to developing a large-scale CPT for gravelly materials. The theoretical and fundamental
understanding of the static push of a CPT is well developed through cavity expansion theory
(see, e.g., Salgado et al., 1997), which can make any required corrections more robust.
Nonetheless, penetration of a large-diameter cone into a gravel stratum may be difficult and
impractical.
Shear wave velocity (Vs) represents an alternative to using large-scale penetration tests to
characterize the liquefaction resistance of gravelly soils. Vs can be measured easily in gravelly
soils, especially when using non-invasive surface wave methods. As Vs is directly correlated to
liquefaction triggering potential, conversion to an equivalent SPT blow count is not required.
Limited measurements of Vs at gravelly sites that liquefied in previous earthquakes have shown
that the Vs of these materials was consistent with existing CRR-Vs relationships developed for
sandy sites (Andrus et al., 1992; Andrus and Stokoe, 2000). More research is needed to confirm
this observation because prior research (Seed and Idriss, 1970; Seed et al., 1986) indicates that
the Vs of gravels is significantly higher than that of sands at the same relative density. Through
further research, however, Vs may be developed as a practical basis to assess the liquefaction of
gravelly materials.

GEOTECHNICAL FIELD DATA FOR LIQUEFACTION TRIGGERING ANALYSES

Site characterization for an analysis of liquefaction triggering includes collection of


information to accurately estimate the values of CRR and earthquake-induced CSR at the site.
Calculation of the CRR requires in situ measurement of the profile of SPT blow count, CPT tip
resistance, or Vs as a function of depth and at multiple locations across the site. Various
correction factors are applied to the measured SPT blow count to define a stress- and energy-
corrected SPT resistance (N1)60 based on the test setup. The energy transferred to the SPT rod
needs to be measured (ASTM D4633) for accurate energy corrections to be applied. Note that the
use of the automatic hammer does not negate the need for energy measurements or the use of a
correction factor. The use of SPT setups that require other adjustments—such as the use of a
non-standard borehole diameter or of samplers configured for liners but used without liners
(even though the setups may be allowed by the ASTM D1586)—need to be avoided.
Furthermore, SPT blow counts recorded in hollow stem auger borings below the water table are
particularly susceptible to error due to soil disturbance and may result in abnormally low blow
count values.
It is important to identify and characterize the detailed variations in the in situ resistance
within a soil profile. The SPT provides measurements at widely spaced intervals (often 1.5 m,
but never less than the length of the split spoon sampler, 0.45 m), which limits the ability to use
SPT measurements to identify thinner layers or detailed variations within a soil profile. Because
the CPT provides almost continuous measurements, it is ideal for characterizing thinner layers
and detailed variations within strata. Depending on the method of measurement, Vs may be used
to identify thin layers and variations within strata, but it does not provide the detail and the
resolution of the CPT. A seismic CPT provides a cost-effective means of measuring Vs while
also providing a detailed profile of in situ tip resistance. Pore-pressure data from piezocone
penetration testing (CPTu) can provide additional information, both qualitative (e.g., whether
soil is dilatant or not) and quantitative (e.g., the steady-state porewater pressure).
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 107

Characterization of soil type is also important to a liquefaction triggering analysis, and CPT and
Vs methods do not provide a direct measure of soil type. In these cases, additional boring and
sampling, or sampling using a special sampler adapted for use with CPT rigs, is needed to
determine soil type directly.
When using liquefaction triggering methods that require Vs values to calculate the
earthquake-induced CSR from site response analysis (see, e.g., Cetin and Seed, 2004), Vs should
be measured directly and not estimated by correlations with the SPT or the CPT. A
comprehensive site investigation for liquefaction triggering on an important project could
include all three characterization techniques: borings with SPT sampling (with hammer energy
measurements—a stricter requirement than use of the automatic hammer) to obtain blow counts
and soil type; CPT soundings to obtain detailed profiles of in situ resistance; and Vs profiles to
accurately assess the earthquake-induced CSR and to provide additional insights into the CRR.

INTERPRETING THE RESULTS OF STRESS-BASED TRIGGERING ANALYSIS

Assessments of liquefaction triggering may not match observed field behavior for many
reasons, including the presence of unsaturated zones beneath the water table, horizontal stresses
due to over -consolidation, or artesian conditions and inadequacies in the simplified methods.
The goal of a liquefaction triggering analysis is to evaluate whether liquefaction is expected to
occur at a site under a given seismic load. Traditionally, this evaluation has been based on
computing the FS, defined as the ratio of CRR to CSR, at different depths within the soil profile.
An FS less than 1.0 is generally assumed to indicate that liquefaction is expected to trigger at that
depth. The FS against liquefaction, however, does not give insights into the associated
uncertainties and variability related to the calculation of CRR and CSR. In practice, a minimum
required FS for design as low as 1.0 has been required when coupled with an extreme ground
motion level (Kavazanjian et al., 2011). Typical minimum values used in practice are between
1.1 and 1.3, but one part the California Code of Regulations calls for an FS of 1.5 under dynamic
conditions, including liquefaction.4
As noted in Box 4.1, different CRR relationships and associated adjustment factors will
produce different values for the FS against liquefaction. These differences may be regarded as a
result of epistemic uncertainty in the liquefaction triggering models due to imperfect data and
imperfect models, which is analogous to the epistemic uncertainty observed in earthquake
ground motion prediction (Douglas, 2010), where different ground motion prediction equations
predict different levels of earthquake shaking for the same magnitude, distance, site conditions,
and other factors. In seismic hazard analysis, it has become standard practice to include this
epistemic uncertainty through the use of a logic tree (Bommer and Scherbaum, 2008; Atkinson et
al., 2014), where each branch of the logic tree represents an alternative model, and weights
assigned to each branch represent the relative degree of belief regarding the accuracy of the
model. This approach has not been widely used to date for liquefaction hazard assessment but
could be used to incorporate the results from the different liquefaction triggering relationships to
account for the epistemic uncertainty. Different liquefaction triggering models can be assigned to
each logic tree branch with weights assigned based on the judgment of the engineer. The FS

4
See http://www.calrecycle.ca.gov/laws/regulations/title27/ch4sb3c.htm#21750.
108 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

against liquefaction is then computed for each model in the logic tree and these values are used
to compute a weighted mean FS using the assigned weights.
Populating a logic tree involves critical evaluation of the available liquefaction models
combined with professional judgment to assign the weights to the logic tree branches. The
critical evaluation includes an assessment of the data used in model development, the techniques
used to derive the applicable adjustment factors and their extrapolation outside the limits of the
data, and the statistical procedures used. All available liquefaction models need not be included
in a logic tree, but clear descriptions and justifications of the developed logic tree are required.
Details regarding the incorporation of uncertainties and expert judgment in probabilistic analyses
can be found in Budnitz and colleagues (1997). Additional research is needed in applying these
approaches to liquefaction analysis, particularly regarding the appropriate logic tree
implementation that will accurately model the epistemic uncertainty across all scenarios of
interest (e.g., fines contents, penetration resistances, depths) considering what is known and
unknown about liquefaction triggering.
The probability of liquefaction is an alternative to the FS. For example, Juang and
colleagues (2002) developed mappings between FS and probability of liquefaction triggering.
Cetin and colleagues (2004) and Idriss and Boulanger (2014) both suggest that the deterministic
triggering curve is associated with a probability of liquefaction of approximately 15%. The
probability of liquefaction has three components: the probability of the earthquake motion used
to compute the CSR, the uncertainty in the relationship between the CRR and in situ test
parameters, and the uncertainty in the parameters that enter into the calculation of CRR and
CSR. Box 4.3 presents a simple analysis that demonstrates that the uncertainty in the adjustment
factors and parameters used in a liquefaction analysis can be larger than the uncertainty in the
location of the CRR curve.
The FS or probability of liquefaction is often computed for a single level of ground shaking
and its associated earthquake magnitude. Over time, though, a given site may experience a wide
range of ground shaking, from frequent weak motions to strong motions that occur only rarely.
Furthermore, each level of ground shaking is associated with a distribution of contributing
earthquake magnitudes, and those distributions vary with the level of shaking. An alternative to
the characterization of liquefaction triggering is a PLHA, in which the seismic hazard curve from
a PSHA is convolved with a probabilistic liquefaction triggering model to compute the mean
annual rate of non-exceedance for different FSs against liquefaction (see, e.g., Kramer and
Mayfield, 2007). Epistemic uncertainty can also be included in a PLHA through a logic tree. The
reciprocal of the mean annual rate of non-exceedance for FS = 1.0 from a PLHA is the return
period of liquefaction (i.e., the average time between occurrences of liquefaction at a site).
Computed in this manner the return period of liquefaction accounts for uncertainties in the local
seismicity, earthquake magnitude, source-to-site distance, and ground shaking as well as the
prediction of liquefaction triggering. It therefore provides a more complete and objective
indication of liquefaction triggering than the FS or probability of liquefaction do based on a
single ground motion value and magnitude. Publicly available computer programs exist that
perform these calculations (see, e.g., Franke et al., 2014; Kramer, 2008), and the calculations can
be incorporated relatively easily within a spreadsheet if the model components are available.
PLHA for liquefaction triggering, as well as its application to the evaluation of liquefaction
consequences, is discussed further in Chapter 9.
The empirical stress-based simplified methods have served the profession well, but they all
are flawed to various degrees. Although a variety of issues associated with their use remains, the
SIMPLIFIED STRESS-BASED LIQUEFACTION TRIGGERING ASSESSMENT 109

technical community will rely heavily on these stress-based methods until reliable and readily
usable mechanics-based approaches are available to evaluate liquefaction and related phenomena
(see Chapter 8). Better results are likely if the practicing engineer employs multiple empirical
methods, the sources of discrepancies between results of the various methods are assessed, and
engineering judgment is applied to resolve those differences. The concepts expressed herein are
reinforced with recommendations provided in Chapter 10.

BOX 4.3
Uncertainty in the Stress-Based Procedures for Liquefaction Evaluation

The factor of safety, FS, against liquefaction triggering in a stress-based analysis is defined as the ratio
between the CRR and CSR, as outlined in Box 1.3. The CSR term at a given depth in the potentially
liquefiable soil is:

𝜎𝑣 𝑃𝐺𝐴 1 (1)
𝐶𝑆𝑅𝑀=7.5 = 0.65 ∙ ′
∙ ∙ 𝑟𝑑 ∙
𝜎𝑣 𝑔 𝑀𝑆𝐹

For the SPT, the CRR is computed from the corrected, clean-sand blow count, (N1)60cs:

(𝑁1 )60𝐶𝑆 = 𝐶𝑁 ∙ 𝐶𝐸 ∙ 𝐶𝐵 ∙ 𝐶𝑅 ∙ 𝐶𝑆 ∙ 𝑁 + ∆(𝑁1 )60 (2)

where CN is the correction factor to adjust the blow count to a reference stress of one atmosphere; CE is a
correction factor for the kinetic energy of the hammer (i.e., hammer weight and height of fall); CB is a
correction factor for the borehole diameter; CR is a rod length correction factor; CS is a correction factor for
the configuration of the SPT sampler; N is the recorded blow count; and (N1)60 is the correction factor for
the fines content (percent passing the #200 sieve by weight) of the soil (Kavazanjian et al., 2011). There is
uncertainty in the computed FS from a stress-based analysis not only because of the uncertainty in the
location of the CRR relationship but also because the values of the parameters in the CSR and (N1)60cs
equations are not known precisely. In fact, explicit consideration of uncertainty associated with a particular
correction factor may even increase the uncertainty associated with the liquefaction potential assessment.
A hypothetical example is used to illustrate the effect of the uncertainty in the input parameters on the
probability of liquefaction. The example is based on Appendix A of Idriss and Boulanger (2008) and has
PGA = 0.28 g, M = 6.9, and a groundwater table depth of 1.8 m. The top 8 m of the profile consist of poorly
graded sand with less than 5% fines. The values of N from the Idriss and Boulanger (2008) profile were
increased to between 12 and 25 to obtain FSs against liquefaction triggering close to 1.5 at all depths.
Estimates of the coefficients of variation (COVs) for the input parameters are provided in Table 1. Assuming
that the uncertain parameters are log-normally distributed, the variance of the logarithm of the product of
several independent, log-normally distributed variables is equal to the sum of the variances of the
logarithmic values. To avoid exaggerating the uncertainty, the values selected for most of the COVs in Table
1 are relatively small. The two exceptions are the COVs of 0.20 for an MSF and rd, which are justified
because of the large range of published values for these parameters.
Figure 1 presents the computed probabilities of liquefaction for the case of perfectly known input
parameters with an uncertain location of the CRR relationship, as well as for the case of uncertain input
parameters with a perfectly known CRR relationship. The uncertain location of the CRR curve is calculated
from equation (6.26) from Idriss and Boulanger (2010). For this example, the probabilities of liquefaction are
very small when the input parameters are certain. The probabilities of liquefaction, however, are about 20%
throughout the layer when accounting for uncertain input parameters, although the analysis assumes that
the CRR relationship is perfectly known. This example shows that the uncertainty in the values of the input
parameters and correction factors is much more important than is the cluster of uncertainty about the mean
of the CRR line in estimating the probability of liquefaction. If the uncertainties associated with spatial
variability in soil liquefaction resistance across a site were included in the analysis, the probabilities could
increase further.
110 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

TABLE 1 COVs for Parameters Used in Liquefaction Triggering Analysis


Description Parameter COV Description Parameter COV
Measured blow count N 0.10 CSR scale factor 0.65 factor 0.05
Each SPT borehole CB, CE, CN, 𝜎𝑣
0.02 Stress ratio 0.05
correction factor CR, and CS 𝜎𝑣′

Magnitude scaling Peak ground


MSF 0.20 PGA 0.10
factor acceleration
Shear stress reduction
CRR stress adjustment Kσ 0.07 rd 0.20
coefficient

FIGURE 1 Probabilities of liquefaction due to the uncertain location of the CRR curve given certain input parameters
and due to perfectly known location of the CRR curve given uncertain input parameters. COURTESY: J. Christian.
5
Alternative Approaches to Liquefaction
Triggering Assessment

Key Findings and Conclusions

 It is important to develop and improve alternatives to the simplified stress-based method for prediction
of liquefaction triggering. Alternatives include strain-based approaches, energy-based approaches,
laboratory and physical model testing, and computational mechanics approaches.
 The relationship between earthquake magnitude and the equivalent number of uniform cycles of
loading in laboratory tests warrants further study. This relationship is important not only in laboratory
assessment of liquefaction triggering but also to the magnitude scaling factor used in the simplified
method.
 Liquefaction hazard maps and geologic screening methods warrant continued development.
 Maps of probabilistic liquefaction hazard could be expanded in coverage, even to that of the national
seismic hazard maps.

There are a growing number of alternatives to the simplified stress-based approach to


liquefaction triggering assessment discussed in Chapter 4. This chapter addresses such
alternative methods that are either proposed or in use to assess liquefaction triggering. One is a
screening tool—regional maps that delineate areas susceptible to liquefaction triggering largely
on geologic grounds. Other methods described include a cyclic strain-based alternative to the
cyclic stress-based approach, energy-based approaches, laboratory and physical model tests, field
measurement of pore-pressure generation under dynamic loading, and computational mechanics
approaches.

REGIONAL LIQUEFACTION HAZARD MAPS AND PAST OCCURRENCE OF


LIQUEFACTION

Geologic maps provide a basis for depicting the potential for liquefaction triggering at a
regional scale. Such depictions are usually employed as planning tools for government agencies

111
112 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

and as guidance for site-specific studies. They are seldom used alone for site-specific
liquefaction evaluations, but they have a role in site characterization (Siegel, 2013) and in
determining whether a detailed liquefaction evaluation is required.
The mapping approach to liquefaction hazard assessment can be traced to a comprehensive
study that correlated mappable geologic units with liquefaction damage during the 1964 Alaska
earthquake (McCulloch and Bonilla, 1970). Later, mappable sedimentary deposits of various
ages and depositional environments were ranked qualitatively by liquefaction potential and used
to develop maps of liquefaction susceptibility (Youd and Perkins, 1978). The past decade has
brought quantitative assessment of liquefaction potential into liquefaction-hazard maps. Modern
liquefaction hazard maps incorporate geotechnical estimates of the liquefaction potential of
individual geologic units based on measured in situ parameters across each unit and probabilistic
estimates of ground shaking (Baise et al., 2006; Holzer et al., 2011). An example of such a map
for Boston, Massachusetts (Brankman and Baise, 2008), is shown in Figure 5.1. Similar maps
have been developed elsewhere: for example, for California (California Geological Survey,
2004, 2008), Indiana and Kentucky (Haase et al., 2011), and South Carolina (Heidari and
Andrus, 2010, 2012).
Regional liquefaction maps could have a role in probabilistic evaluations of liquefaction
triggering. As noted in Chapter 4, the Bayesian methods used to develop several of the
probabilistic triggering relationships assume non-informative prior probabilities of liquefaction.
If the empirical relationships for cyclic resistance ratios (CRRs) developed by Bayesian methods
were presented as CRR curves associated with different likelihood ratios rather than different
probability levels, the curves could be combined with prior probabilities derived from regional
liquefaction maps to obtain more representative posterior probabilities of liquefaction (see Box
4.2).
Regional geologic perspectives on liquefaction hazards also underpin the AASHTO (2014)
guide specifications. If evidence of previous liquefaction is found in the vicinity of a site, then a
detailed liquefaction analysis is always required. Interpreted broadly, the guideline encourages
consideration of evidence for liquefaction not just at the site but also in analogous deposits
nearby, and not just during modern earthquakes but also in the recent geologic past.
Liquefaction maps and geologic screening methods warrant continued development. Maps
of probabilistic liquefaction hazard could be expanded in coverage, even to a national scale
similar to the national seismic hazard maps. The mapping could be anchored in geotechnical data
from the geologic map units. A database of geologic evidence for liquefaction, compiled recently
for the central and eastern United States (Tuttle and Hartleb, 2012), extends the timescale of
observations into previous millennia and could be extended nationwide. Such regional products
would not replace site-specific evaluations; rather, they would provide guidance as to the need
for site-specific evaluations.
ALTERNATIVE APPROACHES LIQUEFACTION TRIGGERING ASSESSMENT 113

FIGURE 5.1 Liquefaction hazard map for Boston, Massachusetts. SOURCE: Brankman and Baise, 2008.

CYCLIC STRAIN-BASED APPROACH

The cyclic strain approach (Dobry et al., 1982) assesses liquefaction triggering through
the prediction of pore-pressure generation as a function of the earthquake-induced shear strain
(see Chapter 2). The approach takes advantage of the fact that cyclic compression
(for drained conditions) and pore-pressure generation (for undrained conditions) are related
more closely to shear strain than shear stress (Silver and Seed, 1971; Youd, 1972; Martin et al.,
1975; Dobry and Ladd, 1980). As shown in Figure 5.2, the relationship between excess pore-
pressure ratio (ru) and cyclic shear strain () for a given number of constant amplitude strain
cycles in strain-controlled tests falls in a relatively narrow band, particularly at low strains
and pore-pressure ratios, across different sands prepared at different relative densities
using different techniques and tested at different effective confining pressures. The
volumetric threshold shear strain, defined as the shear strain below which no pore
pressure develops, is consistently about 0.01% for normally consolidated sands (Dobry and
Abdoun, 2011).
The strain approach to evaluating liquefaction triggering involves predicting the excess
pore-pressure ratio, which requires an estimate of the earthquake-induced shear
strain. The induced shear strain is computed as the earthquake-induced shear stress at a
given depth divided by the shear modulus of the soil at the same depth. The earthquake-
induced shear stress is calculated using the same approach as the stress-based method
(Chapter 4, Equation 4.1), and the shear modulus of the soil is related to the soil shear wave
velocity (Vs) and the reduction in shear modulus with increasing shear strain. This method
ignores the effect of pore-pressure generation on the shear modulus of soil. Given the
computed earthquake-induced shear strain and the number of strain cycles, the pore-
pressure ratio can be estimated from a curve such as shown in Figure 5.2.
114 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 5.2 Relationship between pore-pressure ratio and cyclic shear strain for 10 cycles of loading in strain-
controlled testing. SOURCE: NRC, 1985.

Proponents of the strain-based method argue that the relationship between pore pressure and
shear strain is only moderately influenced by such factors as soil density, soil fabric, and
confining pressure (Dobry et al., 1982; Dobry, 1985), so that a large suite of correction factors is
not necessary. The main parameter used to characterize the in situ soil for the strain approach is
the Vs, which is a more fundamental material parameter than is penetration resistance. Theories
are available that predict how Vs is influenced by effective stress and other variables (Kramer,
1996). Recently, induced shear strain was used as the unifying parameter to interpret liquefaction
triggering charts for in situ parameters and to explain pore-pressure generation from different
centrifuge experiments (Dobry and Abdoun, 2011), which shows the potential for the strain
approach to provide important insights into the prediction of liquefaction triggering.
A limitation of the strain-based approach proposed by Dobry and colleagues (1982), however,
is the computation of the number of equivalent strain cycles needed to use Figure 5.2 or a similar
relationship. The shear strain computed using the procedures outlined above represents the average
shear strain during earthquake shaking. Correlations developed for the number of equivalent stress
cycles could be used (see, e.g., Liu et al., 2001) to evaluate the average shear strain based on an
average shear modulus, but this approach assumes that the equivalency of stress cycles (which is
based on a cyclic strength curve) is the same as the equivalency of strain cycles, and this
assumption is uncertain (Carter and Seed, 1988; Green and Terri, 2005). The issue can be avoided
by using the maximum shear strain after a given number of stress-controlled loading cycles (Cetin
and Bilge, 2012), but in this case, the estimated shear strain includes the effects of pore-pressure
generation, which is not taken into account in other simplified procedures.
One application of the strain approach is to determine if the induced shear strain exceeds
what might be called the “threshold” shear strain (about 0.01%), below which no excess pore
pressure is generated. If the induced shear strain is less than the threshold shear strain, this
method predicts that liquefaction is not of concern. If the induced shear strain exceeds the
threshold shear strain, additional analyses of the potential for triggering of liquefaction and its
consequences are necessary.
Despite its introduction more than 30 years ago, the strain approach has not been a focal
point of further extensive research development, and it has not gained traction in engineering
ALTERNATIVE APPROACHES LIQUEFACTION TRIGGERING ASSESSMENT 115

practice. The uncertainty associated with predicting the strain amplitude may be inhibiting
adoption of strain-based methods. Issues related to the relationship between the number of
equivalent strain cycles and earthquake magnitude, many of which are similar to issues related to
the number of equivalent stress cycles discussed subsequently in the section on laboratory
assessment of liquefaction triggering, must also be resolved with respect to the strain approach.
In addition, full validation and calibration of the strain approach against field case histories is
needed for this approach to become a more accepted method of assessing the potential for
triggering liquefaction. Research in this area would be of value.

ENERGY-BASED APPROACHES

Liquefaction triggering approaches that employ ground motion intensity measures other
than the peak ground acceleration (PGA) commonly have been classified as “energy-based”
approaches, regardless of whether the intensity measure is truly a measure of energy. Many of
the proposed procedures were developed using an approach similar to that used for the simplified
stress-based procedures (i.e., correlations relating in situ test measurements to the liquefaction
resistance derived from field case histories).
Input energy is fundamental to the breakdown of the soil fabric, which, in turn, leads to
pore-pressure generation and liquefaction. Seismologists quantify the energy released during
earthquakes with simple equations relating seismic wave energy to earthquake magnitude (see,
e.g., Gutenberg and Richter, 1956). Nemat-Nasser and Shokooh (1979) developed a mechanics-
based relationship between dissipated energy and excess pore-pressure generation, and several
experimental studies developed similar empirical relationships (e.g., Davis and Berrill, 1982;
Berrill and Davis, 1985; Yamazaki et al., 1985; Yanagisawa and Sugano, 1994; Liang, 1995;
Green et al., 2000, Polito et al. 2008). As a result, it has been concluded that research and
development of energy-based approaches should continue (Youd et al., 2001).
The proposed energy-based approaches have several forms, but they are all founded on
quantifying both the seismic demand and the liquefaction resistance in terms of energy. One of
the earliest energy approaches was based on dissipated energy, which is proportional to the
cumulative area bounded by stress-strain hysteresis loops (Figure 5.3a). One such cumulative
energy approach for evaluating the potential for liquefaction triggering was proposed by Davis
and Berrill (1982) and modified by Berrill and Davis (1985). The seismic demand is computed
from the estimated seismic wave energy released at the earthquake source (Gutenberg and
Richter, 1956), modified to account for the attenuation of the energy with distance due to both
geometrical spreading (Davis and Berrill, 1982) and material damping (Berrill and Davis, 1985).
These investigators coupled this modified seismic wave energy with an empirical relationship
that relates dissipated energy per unit volume of soil to excess pore-pressure generation to
develop an expression for seismic demand. Using case history data, a correlation was developed
between corrected standard penetration test (SPT) N-values and the dissipated energy required to
cause liquefaction (see Figure 5.3b). The biggest drawbacks of the approach are the relatively
large variability in the estimated seismic energy demand and the lack of a unique link between
seismic wave energy arriving at the site and dissipated energy in a stratum at a given depth.
Nevertheless, their approach was pioneering and set the stage for several subsequent studies that
apply more refined approaches to estimate dissipated energy in a soil profile (e.g., Law et al.,
1990; Green, 2001; Kokusho, 2013).
116 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

(a) (b)

FIGURE 5.3 (a) Schematic diagram of dissipated energy per unit volume of soil in a cyclic triaxial test specimen. The
dissipated energy is proportional to the cumulative area bound by the stress-strain hysteresis loops. SOURCE:
Yoshimoto, N., R. Orense, M. Hyodo, and Y. Nakata. 2014. Dynamic behavior of granulated coal ash during
earthquakes. Journal of Geotechnical and Geoenvironmental Engineering 140(2). With permission from ASCE. This
material may be downloaded for personal use only. Any other use requires prior permission of the American Society
for Civil Engineers. This material may be found at [http://ascelibrary.org/doi/abs/10.1061/%28ASCE
%29GT.1943-5606.0000986]. (b) Relationship between SPT blow count and dissipated energy required to
cause liquefaction proposed by Berrill and Davis (1985). SOURCE: Berrill and Davis, 1985. (r = closest
distance to earthquake fault rupture, v = vertical effective stress, M = earthquake magnitude, and A = Material
Attenuation factor).

Proponents of energy-based procedures state that these procedures have the potential to
capitalize on the strengths of both the stress-based and the strain-based procedures because
energy per unit volume of soil is a function of both stress and strain (Green, 2001).
Other reported advantages of the energy-based approaches are the absence of a magnitude
scaling factor (MSF) and the ability to account for differences in the characteristics of
shaking in different tectonic regions (e.g., Central and Eastern North America relative to
Western North America) and for non-earthquake sources of vibration (see, e.g., Green and
Mitchell, 2004). Energy-based procedures are also subject to some of the limitation of other
procedures, however. For example, for some of the proposed procedures, the effects of pore-
pressure generation are ignored when estimating the dissipated energy demand, similar to
the approach taken in the stress-based and strain-based methods. Although not yet used
in practice, research and development of energy-based approaches should continue because
the potential advantages of the approach over currently used procedures appear promising.
Procedures that use Arias intensity (Arias, 1970) as a measure of the strength of earthquake
shaking have also been referred to as energy-based procedures (see, e.g., Kayen and
Mitchell, 1997). Originally developed to quantify the damage potential of earthquake
motions on buildings, Arias intensity is proportional to the integral of the acceleration-time
history squared over time. As a result, it is influenced by the amplitude variation,
frequency content, and duration of the earthquake motion. As such, Arias intensity, then,
could be considered a reasonable measure to quantify seismic demand, and several
ground motion predictive relationships have been developed to estimate Arias intensity
as a function of earthquake magnitude, distance, and other parameters (see, e.g., Kayen and
Mitchell, 1997; Travasarou et al., 2003; Lee and Green, 2008, 2010). Nevertheless, the use
of Arias intensity as the sole
ALTERNATIVE APPROACHES LIQUEFACTION TRIGGERING ASSESSMENT 117

intensity measure for quantifying seismic demand for evaluating liquefaction triggering potential
has been questioned (see, e.g., Green and Mitchell, 2003). Specifically, laboratory studies have
shown that liquefaction resistance is independent of the frequency of the applied loading for the
range of frequencies of interest in earthquake engineering (Riemer et al., 1994), but Arias
intensity is influenced by the frequency content of an earthquake motion. Furthermore, the large
standard error in Arias intensity ground motion prediction equations suggests that there will be
large uncertainty in the prediction of seismically generated pore pressures based on Arias
intensity even if a unique relationship between pore-pressure generation and Arias intensity is
established.

LABORATORY AND PHYSICAL MODEL TESTING

Within the context of this report, laboratory testing refers either to index property testing for
soil characterization purposes or to single-element cyclic laboratory tests in which a presumably
representative homogeneous specimen is subject to known stress or strain boundary conditions
and is cyclically loaded. Index property tests are used to classify the soil and to assess the
susceptibility of a soil to liquefaction triggering either directly (e.g., on the basis of grain size
and plasticity) or indirectly (e.g., by validating soil classification obtained indirectly from cone
penetration test [CPT] data). Grain size and plasticity data may indicate that the potential for
liquefaction triggering can be assessed using the simplified method (with percent passing the
#200 sieve as an input parameter); that the soil is not susceptible to liquefaction triggering (as
discussed in Chapter 4); or that the soil is amenable to sampling and cyclic laboratory testing for
evaluation of the potential for liquefaction triggering (as discussed in this chapter).
In addition to grain size and soil plasticity, the maximum and minimum density (or unit
weight) of a granular soil may also be useful parameters for establishing the relative density1 of
soil used in laboratory testing, as liquefaction triggering potential of a granular soil in situ is
generally considered to be related to relative density and not the absolute density of the soil. In
situ characterization of relative density usually relies on indirect measurement via correlation
with SPT or CPT resistance given the difficulty in recovering relatively undisturbed samples. In
situ relative density evaluated on the basis of CPT or SPT resistance, however, is rarely, if ever,
used to assess liquefaction triggering potential. Instead, CPT or SPT resistance is used directly to
assess liquefaction triggering potential (e.g., using the simplified method).
In this report, the term “physical model testing” refers to either “1-g”2 shake table testing or
dynamic centrifuge model testing. The designation of “1-g” assigned to shake table testing
distinguishes it from centrifuge testing wherein models are loaded via a shake table but the shake
table is mounted in a geotechnical centrifuge and centrifugal acceleration is used to multiply the
body forces on the soil skeleton (for model scaling purposes). Physical model testing may be
used to evaluate the behavior of potentially liquefiable systems subject to earthquake loading,
including layered soil profiles and soil-structure systems. Layered soil profiles may include
horizontally layered stacks of soil (even if the soil type is uniform—e.g., to evaluate confining

1
The relative density of a soil is equal to (emax-e)/(emax-emin) where e is the void ratio of the soil, emax is the maximum
void ratio of the soil (representative of its loosest state), and e min is the minimum void ratio of the soil (representative
of its densest state).
2
g is the gravitational field equal to 9.8 Newtons per kilogram.
118 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

pressure effects and to validate numerical models), sloping soil profiles, and profiles with
embankments or a free face (e.g., a riverbank). Due to the dependence of liquefaction triggering
potential on the stress level in the soil, and given the difficulty and expense of large-scale 1-g
shake table testing, the usefulness of 1-g shake table tests is limited. Centrifuge model testing
allows scaled model testing, with the model scale equal to the centrifuge acceleration (in
multiples of the acceleration of gravity). Modern geotechnical centrifuge model testing began in
the 1970s and matured since 1985 as a tool to investigate factors that influence liquefaction
triggering and consequences. Nevertheless, the efficacy of centrifuge model tests for liquefaction
studies has been questioned given sample preparation issues, grain size and permeability effects,
and the influence of boundary conditions unique to centrifuge models.
There have been few advances in laboratory testing for liquefaction triggering assessment
since the 1985 report (NRC, 1985) on liquefaction was published. This, perhaps, reflects both the
great difficulty in obtaining high-quality samples (samples in which the soil structure is
relatively undisturbed) for testing and the trend toward in situ testing to evaluate soil liquefaction
resistance. Nonetheless, laboratory testing remains a valuable tool for investigating the influence
of factors such as effective overburden pressure and fines content on liquefaction triggering
potential and for assessing triggering potential in soils amenable to recovery of high-quality,
relatively undisturbed samples. For example, soils with significant fines content, particularly if
the fines are plastic, may be amenable to laboratory testing to determine their in situ resistance to
triggering. Furthermore, the applicability of the simplified method to assessing the liquefaction
triggering potential of these types of soils is suspect, as noted in Chapter 4. It is therefore often
desirable to recover relatively undisturbed high-quality samples of potentially liquefiable finer-
grained soils using thin-walled tubes, preferably using special techniques such as fixed piston
sampling (Osterberg, 1952; ASTM D6519) or gel push sampling (Huang et al., 2008). Reimer
(2007) describes laboratory testing of low-plasticity silty soil to evaluate its liquefaction
triggering potential for the Bay Area Rapid Transit expansion project from Fremont to San Jose,
California.

Cyclic Laboratory Testing

Laboratory tests that can be used to evaluate or provide insight into liquefaction triggering
in soils include cyclic triaxial (CTX) tests and cyclic direct simple shear (CyDSS) tests (Kramer,
1996). CTX tests use a cylindrical sample with a height-to-diameter ratio typically equal to 2.0
or greater encased by a flexible membrane and rigid caps on either end, as illustrated in Figure
5.4. The radial boundary condition in CTX tests is generally stress controlled (i.e., through
application of a constant confining pressure in the cell that encloses the specimen). Cyclic
loading in a CTX test is usually applied by pushing and pulling on the piston attached to the top
cap. CyDSS tests use cylindrical specimens with a height-to-diameter ratio typically equal to 0.5.
The radial boundary condition in this test is typically strain controlled, with the specimen
confined against lateral deformation by a rigid stack of sliding rings, as illustrated in Figure 5.5,
or a wire-wrapped membrane. Both approaches are intended to enforce a zero lateral strain
boundary condition representative of vertically propagating shear waves in a horizontally layered
soil profile. Cyclic loading in a CyDSS test is usually applied via application of a shear force or
shear displacement on a horizontal plane to either the top or the bottom of the specimen (with the
other end fixed), as illustrated in Figure 5.5, and it is assumed that a complementary shear stress
ALTERNATIVE APPROACHES LIQUEFACTION TRIGGERING ASSESSMENT 119

(necessary for equilibrium of the specimen) develops on vertical planes along the sides of the
specimen.

FIGURE 5.4 Typical triaxial test configuration. The cell fluid is pressurized to apply a confining pressure to the soil
specimen, and the cyclic load is applied by pulling and pushing the piston up and down. SOURCE: Modified from
CONTROLS GROUP.

FIGURE 5.5 Cyclic simple shear loading of a soil specimen. Annular brass rings (inside diameter equal to the
diameter of the top cap) enforce a zero lateral strain condition representative of vertically propagating shear waves.
SOURCE: W. Farrance, CONTROLS GROUP.
120 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Cyclic hollow cylinder torsional shear tests can also be used to assess soil liquefaction
triggering in the laboratory. In these tests, a vertically oriented hollow cylindrical specimen
encased in membranes on the inside and outside of the hollow cylinder is mounted in a cell
between rigid base and top caps, subject to a constant confining pressure, and loaded cyclically
by a torsional force of displacement applied to either the top or the base. Hollow cylindrical
specimens must be used to maintain a relatively constant shear strain on a horizontal plane in the
specimen, as shear strain varies linearly from zero at the center of a torsionally loaded cylinder to
a maximum value along the outer perimeter. Due to difficulties in preparing a hollow cylindrical
specimen of cohesionless soil, however, this type of test is much less commonly used for
investigating liquefaction than CTX or CyDSS tests.
CTX tests are easier to perform than CyDSS tests, but CyDSS tests provide a better
representation of both the in situ initial (K0) stress and strain state in level ground and the
earthquake-induced stresses due to vertically propagating shear waves. As illustrated in Figure
5.6, simple shear loading accurately reflects the initial (at rest) in situ stresses in level ground
(i.e., the K0 stress state) and the stresses imposed on an element of soil by a vertically
propagating shear wave in horizontally layered soil (i.e., imposition of a cyclic shear stress with
no change in normal stress on the vertical and horizontal planes). The CTX test applies the cyclic
shear stress that represents the earthquake load on the 45o plane of the laboratory specimen—i.e.,
not on the horizontal plane to which the earthquake shear stress is applied in the field. The
normal stress on that 45o plane is not constant as it is on horizontal plane in the field, but instead
varies about the mean normal stress.
The stress state in the hollow cylinder torsion test also closely approximates the stresses
induced in the ground by a vertically propagating shear wave, since the vertical and horizontal
normal stresses are constant and a shear stress is applied to the horizontal plane (Towhata and
Ishihara, 1985). But this test is typically limited to testing of reconstituted specimens and is
rarely, if ever, used to test undisturbed samples due to the difficulty in trimming an undisturbed
cylindrical sample and mounting it in the device. CyDSS, CTX, and cyclic hollow cylinder
torsional shear tests can all be used to study patterns of behavior that affect the liquefaction
triggering potential of soils in situ (e.g., the effect of the initial effective stress level and the
effect of an initial static shear stress) on reconstituted samples.
Cyclic laboratory test methods (CTX, CyDSS, and hollow cylinder torsional shear) can
provide substantial insight into the factors that affect liquefaction triggering. Laboratory cyclic
tests have had and will continue to have an essential role in investigating and evaluating, for
example, the influence of density (or relative density), fines content, particle size and size
distribution, degree of saturation, and nonuniform load cycles on the potential for liquefaction
triggering. As noted in Chapter 4, the extrapolation of the relationships used in the simplified
method for overburden stress, grain size, and magnitude effects into ranges where they are not
constrained by field data need to be informed by this type of laboratory testing. CTX testing
remains popular for this purpose due to the ease with which the tests can be performed and the
availability of equipment. Caution is warranted, though, when using CTX tests due to the
difference in stress state between the CTX test and the field condition. Laboratory testing is also
useful for developing constitutive relationships for pore-pressure development prior to
liquefaction and for the post-liquefaction behavior of soils.
It is difficult to obtain relatively undisturbed samples of cohesionless liquefiable soil. As
noted previously, soils with enough plastic fines can maintain their structures if sampled using
specialized thin-walled tubes and sampling equipment that minimizes disturbance. Recovering
ALTERNATIVE APPROACHES LIQUEFACTION TRIGGERING ASSESSMENT 121

samples of sandy or gravelly soils without enough fines to maintain their structure during
sampling, extrusion, and mounting in the test cell requires techniques such as ground freezing.
Use of ground freezing to recover relatively undisturbed samples for laboratory evaluation of
liquefaction triggering potential relies on the assumption that a unidirectional freezing front will
push porewater that expands as it freezes out ahead of the freezing front without disturbing the
soil, thus preserving the soil structure and maintaining its properties (Yoshimi et al., 1978;
Hofmann et al., 2000). Note, however, that ice continues to expand in volume during freezing,
and therefore the assumption that no soil volume change is induced by a unidirectional freezing
front is questionable. The question is raised, then, of whether freezing the soil for sampling and
then thawing the soil for testing disrupts the fabric or erases any aging effects that may have
occurred in the field, even given zero net volume change of the soil skeleton during this process.

(a) (b)
FIGURE 5.6 (a) Stresses due to vertically propagating shear waves. SOURCE: Kavazanjian et al., 2011. Note that
the figure makes no representation as to the sign of shear stress. (b) Stresses in cyclic laboratory tests. COURTESY:
E. Kavazanjian.

Ignoring any loss of age-related increase in liquefaction resistance by fabric changes


associated with freezing and thawing would be conservative. If the geologic age of the deposit is
known, the effect of aging could be accounted for by empirically adjusting the laboratory-
measured resistance based on the in situ Vs using concepts developed by Andrus and colleagues
(2009). Another technique that may effectively preserve the physical structure of the
specimen
122 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

prior to sampling is impregnating the ground with a biopolymer (Sutterer et al., 1996). This
approach is burdensome, however, and has been investigated only a few times for research
purposes. Due to the expense and difficulty in obtaining relatively undisturbed samples through
ground freezing or other special techniques, laboratory testing for site-specific evaluation of
liquefaction triggering in sandy and gravelly soils is generally limited to projects of high value,
where there are critical consequences, and for research purposes. A technique that allows cost-
effective recovery and testing of cohesionless soil samples such that in situ cyclic resistance was
retained would facilitate laboratory testing for liquefaction triggering evaluation of such soils in
situ for more routine projects.

Assessment of Liquefaction Triggering Potential from Cyclic Laboratory Tests

While nonuniform load sequences can be applied in cyclic laboratory testing, including load
sequences based on strong motion time histories, there is no generally accepted procedure for
performing and interpreting the results of such laboratory tests. Therefore, most laboratory test
programs employ uniform cycles of loading. In a typical laboratory-based liquefaction triggering
assessment, the objective is to evaluate the cyclic strength curve for the soil (see Chapter 2,
Figure 2.7). If only a limited number of cyclic tests can be performed, they can be used to anchor
a typical cyclic strength curve from the literature for a soil of similar composition. Once the
cyclic strength curve is known, the factor of safety can then be assessed based on the
representative cyclic stress ratio (CSR) from the simplified stress-based approach (see Chapter 4,
Equation 4.2) and Neq, the representative number of uniform cycles for the design earthquake.
Alternatively, a laboratory-based assessment of liquefaction triggering potential can be
accomplished by applying the representative CSR to the specimen and comparing the number of
cycles of loading to liquefaction under this stress ratio to Neq.
If enough specimens are available, laboratory testing can be used to generate the complete
cyclic strength curve. The cyclic strength curve still must be related to the specific value of the
CRR, however, to use in the liquefaction potential assessment based on Neq. Thus, choosing the
correct value of Neq is a key issue in a laboratory-based evaluation of liquefaction triggering.
Additionally, it must be kept in mind that laboratory tests evaluate only the liquefaction on an
element level (i.e., on the sample being tested) as opposed to the larger liquefiable soil mass
present in the field.
Seed and colleagues (1975b) used a variant of the Palmgren-Miner (P-M) high-cycle fatigue
theory (Palmgren, 1924; Miner, 1945) to develop a correlation between Neq and earthquake
magnitude from approximately 60 strong motion recordings from North and South America. The
Seed and colleagues (1975b) approach to compute Neq consists of taking a weighted average of
the “peaks” in an acceleration-time history, wherein the amplitude of a peak was defined as the
maximum value of the time history between successive zero crossings (known as the zero-
crossing method; ASTM, 2011). Seed and colleagues (1975b) based their weighting function on
the shape of the laboratory cyclic strength curve shown in Figure 5.7 and used 0.65 amax, where
amax is the peak acceleration in the earthquake record, as their reference value for the amplitude
of the peaks in the time history: that is, peaks having an amplitude of 0.65 amax were given a
weight of one; peaks having amplitudes greater than 0.65 amax were given weights greater than
one; and peaks having amplitudes greater than 0.65 amax were given weights less than one. The
ALTERNATIVE APPROACHES LIQUEFACTION TRIGGERING ASSESSMENT 123

resulting relationship between Neq and earthquake magnitude is shown in Figure 5.7. The Neq
values in Figure 5.7 have remained the standard of practice for over 20 years (Youd et al., 2001).
Neq is intimately related to the MSF used in the simplified method for liquefaction
assessment described in Chapter 4. Seed and Idriss (1982) derived their MSF values from Figure
5.7 by dividing the CSR at Neq for magnitude (M) by the CSR at Neq equal to 15, the value for
Neq for M = 7.5 (the reference magnitude). More recent studies on Neq (and the MSF) rely on the
observation that when plotted on log-log axes, the laboratory cyclic strength curve generally
plots as a straight line with a slope of -b, as illustrated in the Figure 5.8, and a log-log plot of
CSR divided by MSF versus Neq will therefore be a horizontal line (i.e., straight line with the
slope equal to zero). The relationship between Neq and MSF is solely a function of the “b-value”
in Figure 5.8 (see, e.g., Liu et al., 2001; Green and Terri, 2005; Boulanger and Idriss, 2015). If b
and Neq are known, the corresponding MSF is easily derived. Alternatively, Neq values can be
derived from the MSF values. The committee is not aware, however, that reconciliation of Neq
values with MSF values derived from field data has not been attempted.

FIGURE 5.7 Representative relationship between earthquake magnitude and the number of uniform cycles of loading
to cause liquefaction used to develop the Seed and colleagues (1975b) magnitude scaling factors. SOURCE: Seed,
H.B., and I.M. Idriss. 1982. Ground motions and soil liquefaction during earthquakes. Oakland, CA: Earthquake
Engineering Research Institute. With permission from the Earthquake Engineering Research Institute.

FIGURE 5.8 Relationship among laboratory cyclic strength curve, N eq (assumed to equal Nliq), and MSF. Dividing
CSR by MSF results in the cyclic strength curve becoming horizontal. For historical reasons, M w7.5 is used as the
reference magnitude for MSF (i.e., MSF = 1 for Mw7.5). SOURCE: Green, 2001.
124 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Studies on Neq since 2001 have included defining peaks in an acceleration-time history
using alternative approaches: namely, approaches other than the zero-crossing method used by
Seed and colleagues (1975b) (see, e.g., Hancock and Bommer, 2005; Stafford and Bommer,
2009); accounting for the effect of multidirectional shaking (see, e.g., Liu et al., 2001; Carter et
al., 2014); and accounting for the influence of the soil’s stress-strain response (see, e.g., Green
and Terri, 2005; Boulanger and Idriss, 2015). Important conclusions drawn from these studies
include that Neq is negatively correlated with the peak ground acceleration (see, e.g., Biondi et
al., 2004); that Neq is a function of site-to-source distance (see, e.g., Liu et al., 2001; Green and
Terri, 2005; Stafford and Bommer, 2009); and that Neq may vary with depth in the soil profile
(see, e.g., Green and Terri, 2005). Green and Terri (2005) questioned the appropriateness of
using a high-cycle fatigue theory for computing Neq for liquefaction evaluations. These findings
have profound implications, not only for laboratory assessment of liquefaction potential but also
for the simplified method itself due to the intimate relationship of Neq and the MSF. If Neq
depends on, among other factors, site-to-source distance, peak ground acceleration, response
characteristics of the soil, and depth in the profile, then the MSF may also depend on these
variables. In this vein, Boulanger and Idriss (2015) proposed MSF values that are a function of
soil response characteristics, quantified in terms of the slope (i.e., b-value) of the laboratory
cyclic strength curve (see Figure 5.8) and correlated to in situ penetration resistance.
At the present time, there is no consensus on either Neq or MSF values for use in
liquefaction triggering analyses. Considering the importance of Neq to both laboratory
assessment of liquefaction triggering potential and to the MSF used in the simplified method for
liquefaction triggering assessment and their interrelationship, this issue deserves further
attention.
Approaches for site-specific liquefaction assessment that employ laboratory testing also
include a hybrid approach in which the laboratory test results are used to calibrate a numerical
pore-pressure generation model and the numerical model is then used to assess the liquefaction
triggering potential of the soil. Such approaches are not well established, but they can provide
useful information on major projects where the additional expense is warranted. Cyclic
laboratory testing can also be used to determine if a fine-grained soil exhibits what Boulanger
and Idriss (2006) described as “sand-like behavior,” and thus are susceptible to liquefaction
triggering, or to what they described as “clay-like behavior,” and thus are susceptible to cyclic
softening but not to liquefaction.
While laboratory testing to identify factors that influence liquefaction triggering potential
and to develop constitutive models can be conducted using reconstituted specimens, the use of
reconstituted specimens to establish the liquefaction triggering potential of soil in situ is
problematic given difficulties establishing whether the fabric of the soil in situ is reproduced
properly. Mulilis and colleagues (1977) showed that samples reconstituted to the same density
using different compaction techniques can have cyclic shear strengths that vary by more than a
factor of two due to differences in the soil fabric. Some studies, however, have suggested that
small strain Vs may be used as an index for soil fabric. Specifically, samples of a given soil
reconstituted by various methods to the same Vs have been shown to have similar cyclic
resistance (Tokimatsu et al., 1986). This finding has led some investigators to reconstitute
samples in the laboratory to the same Vs that was measured in situ, rather than to the same
relative density (see, e.g., Wang et al., 2006; Baxter et al., 2008), assuming that the cyclic
resistance of the reconstituted samples will then be similar to that of the undisturbed samples.
There is no rigorous proof of this assumption, however, and the lack of confidence in the ability
ALTERNATIVE APPROACHES LIQUEFACTION TRIGGERING ASSESSMENT 125

to replicate in situ fabric in reconstituted samples remains a major obstacle for laboratory-based
approaches to evaluating liquefaction triggering potential. A demonstration that specimens
reconstituted to their in situ shear wave velocity have the same cyclic resistance as the in situ soil
would likely result in increased use of laboratory testing for evaluating the liquefaction
triggering potential of soil in situ.

Physical Model Testing

As described earlier, physical model tests that can be used to study liquefaction triggering
include 1-g shake table tests and centrifuge shake table tests. In a shake table test, a soil deposit
is created on a platform connected to an actuator that shakes the platform. The actuator typically
can be programmed to apply periodic (e.g., sinusoidal), random, or earthquake-like cyclic loads.
Most shake tables translate in only one horizontal direction, although some can apply cyclic
loads in two orthogonal horizontal directions. Certain shake tables apply vertical, rocking, and
torsional loads, but these are generally used to evaluate structural response rather than
geotechnical site response (e.g., liquefaction).
A 1-g model relies on gravity to apply body forces to the model. Therefore, soil in a 0.3-
meter-tall model on a 1-g shake table behaves exactly like 0.3 meters of soil. Because soil
behavior depends in a nonlinear manner on effective overburden stress, the properties and
behavior of 0.3 meters of soil cannot be extrapolated to behavior of soil at greater depth. In
geotechnical centrifuge testing, the physical model is mounted on a platform (sometimes referred
to as a bucket) on the end of an arm that rotates horizontally. The axis on which the bucket
swings is also horizontal and oriented perpendicular to the arm. As the arm rotates, the bucket
“swings up” such that the body force vector (the resultant of the vertical gravity vector and the
horizontal centrifugal acceleration vector) is always perpendicular to the platform. When the
rotating arm reaches its terminal velocity—and therefore its terminal centrifugal acceleration—a
shake table attached to the platform loads the platform in the same manner as a 1-g shake table
loads its platform (i.e., cyclically, in a direction parallel to the base of the platform).
Centrifuge testing is usually conducted at centrifugal accelerations of 20 g or more, resulting
in the bucket swinging up into a nearly horizontal position. The benefit of centrifuge model
testing is that the body forces in the soil, and thus the length scale of the model, are related to the
prototype (i.e., field) values by a factor equal to the resultant of gravity and the centrifugal
acceleration (Kutter, 1992). Hence, theoretically, a 0.3-meter-tall layer of soil accelerated to 50 g
in a centrifuge will have a model-to-prototype length scale of approximately 50 and thus behave
like a 15-meter layer (50 × 0.3 meter) of soil. Such issues as particle size and permeability
scaling, however, as well as the boundary conditions in the centrifuge container, can cloud the
results of a centrifuge model test.
Both 1-g and centrifuge shake table testing of soil deposits are constrained by the lateral
boundary condition in the soil container. Rigid wall containers result in all points along the
vertical boundaries of the soil container moving in exactly the same manner horizontally (i.e.,
not allowing for any relative displacement of horizontal soil layers around the perimeter of the
box—unlike a real soil deposit). The rigid walls of the container may inhibit the development of
shear stress on the horizontal plane, inhibiting the liquefaction of the soil. The rigid walls may
also create multiple reflections of the simulated earthquake waves off the sides of the soil
container, resulting in complex stress states that do not represent field stress states. Therefore,
126 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

both types of shake table tests (1-g and centrifuge) typically employ a laminar box soil container
(i.e., a soil container made of a stack of thin, rigid horizontally oriented frames linked by a
flexible membrane). The membrane allows them to move relative to each other in a plane
parallel to the base of the container (i.e., parallel to the direction of shaking).
As noted above, 1-g model tests are limited by the dependence of liquefaction triggering on
the magnitude of the effective overburden stress and the limited magnitude of the effective
overburden stress achievable in 1-g models. Findings on liquefaction triggering behavior from a
0.3-meter-tall 1-g model simply cannot be used to infer soil behavior at depths greater than 0.3 m
due to the many aspects of soil behavior that depend, in a nonlinear manner, on the effective
overburden stress (e.g., strength, stiffness, compressibility, and shear-induced volume change).
To be useful for the study of liquefaction triggering phenomena, 1-g shaking tables must be large
enough that the stress levels in the soil approach relevant stress levels in the field (i.e., stress
levels associated with depths on the order of at least several meters).
A few 1-g shake tables can accommodate models of the requisite size to study liquefaction
triggering phenomena. Both the E-Defense shake table in Japan3 and the shake table at the
University of California, San Diego4 can accommodate models approaching 10 meters in depth.
Only the E-Defense table, however, can test saturated soil deposits for liquefaction studies. The
earthquake engineering laboratory at the State University of New York (SUNY) at Buffalo
developed a 6-meter-tall laminar box for liquefaction studies under the National Science
Foundation’s Network for Earthquake Engineering Simulation (NEES) program.5 While the box
has experienced experimental difficulties during development, it is, as of this writing, in a
serviceable condition (personal communication, S. Thevanayagam, SUNY at Buffalo, July 19,
2016).
1-g shake table tests may also yield information useful for validating computational models
and studying soil-structure interaction mechanisms and patterns of behavior of liquefied soil
(e.g., the behavior of pile groups behind quay walls) (Motamed and Towhata, 2010). 1-g shake
table tests have also proven useful for investigating the influence of multidirectional shaking and
sample preparation effects on liquefaction triggering (Pyke et al., 1974; De Alba et al., 1975;
Mulilis et al., 1975). These types of test, however, are subject to the limitations on effective
overburden stress discussed above. Even if an appropriate large-scale 1-g shake table is
available, the time and cost associated with large-scale testing is high: for instance, it can take
over a month and several tens of thousand dollars to prepare and conduct one test. The cost of a
series of multiple tests for systematic evaluation of the influence of effects such as particle
gradation, relative density, or earthquake magnitude would likely be prohibitive. The use of
large-scale 1-g shake table tests for site-specific evaluation of liquefaction triggering is also
problematic given issues associated with preparation of a large-scale soil deposit in the model
that is representative of the soil in situ (as discussed above with respect to preparation of
specimens for laboratory testing).
If it can be demonstrated that the results of centrifuge tests in which liquefaction is triggered
are reproducible and representative of field behavior, centrifuge model testing can provide a
means of cost-effective testing for systematic evaluation of liquefaction triggering effects and for

3
See http://www.bosai.go.jp/hyogo/ehyogo/index.html.
4
See http://nheri.ucsd.edu.
5
See http://nees.buffalo.edu.
ALTERNATIVE APPROACHES LIQUEFACTION TRIGGERING ASSESSMENT 127

validating computational models for liquefaction triggering and its consequences. While not
inexpensive, the cost of a series of centrifuge tests would typically be an order of magnitude less
than the cost of a similar series of large-scale 1-g model tests. Furthermore, large capacity
geotechnical centrifuges like that at the University of California at Davis facility can accelerate
models with soil layers on the order of 1 meter tall to centrifuge accelerations exceeding 60 g,
corresponding to a 60-meter-deep soil deposit. Nevertheless, reliance on centrifuge model testing
to study liquefaction phenomena has been hindered by a lack of standardized protocols for
preparing experimental models and conducting tests and by the inability to reproduce test results
at different test facilities (Arulanandan and Scott, 1993).
Advances in centrifuge testing have led to a better understanding of the factors that
influence centrifuge test results. The international Liquefaction Experimentation and Analysis
Protocol (LEAP) program has been established to demonstrate that centrifuge model testing can
give consistent and reliable results for the study of liquefaction phenomena (Manzari et al.,
2014). Issues to be resolved include the influence of model preparation techniques on results;
how to address permeability and grain size scaling issues (sand particles in a centrifuge may look
like boulders at the prototype scale, and a viscous pore fluid must be used in centrifuge testing
instead of water to satisfy permeability scaling); and the influence of boundary conditions
associated with the centrifuge soil container (e.g., use of a laminar box versus boxes with hinged
walls and other types of containers).
If centrifuge testing can be shown to be reliable for the study of liquefaction triggering, it
can provide a means to evaluate many factors that influence triggering, including the same
factors investigated with other types of laboratory testing (e.g., the influence of density or
relative density, grain size distribution [including fines content], and initial effective stress).
Centrifuge testing can also be used to investigate, for example, the influence of nonuniform soil
profiles, sloping ground, soil non-homogeneity, and earthquake time histories (e.g., magnitude,
duration, the sequence of nonuniform load cycles). Understanding the influence of these factors
on liquefaction response can be useful not only for developing methods to predict liquefaction
triggering but also for interpreting field data if there is better understanding of the factors that
affect the complex response of a field site. With greater reliability of centrifuge testing for
predicting liquefaction triggering there can be greater confidence assessing the consequences of
liquefaction (Chapter 7) and for validation of computational models (Chapter 8).

FIELD MEASUREMENT OF PORE-PRESSURE GENERATION


UNDER DYNAMIC LOADING

In situ measurement of pore-pressure generation during dynamic loading is another


alternative for evaluating liquefaction triggering. This approach requires a controlled dynamic
loading source and embedded sensors that measure the dynamic loading and the generation of
pore pressures in the ground. The two most common sources of dynamic loading are controlled
blasting (see, e.g., Charlie et al., 1992; Gohl et al., 2001) and a dynamic vibroseis truck (see, e.g.,
Rathje et al., 2005; Cox et al., 2009).
Controlled blasting has been used for many years to densify soils, but more recently it has
been used as a research tool to induce large excess pore pressures within liquefiable soil deposits
to study the response of various in situ structures within the liquefied soil (see, e.g., Ashford and
Rollins, 2002; Ashford et al., 2006). Because blasting loads the soil rapidly and induces extreme
128 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

ground accelerations, it is difficult to relate the blasting load level to an earthquake loading level
using a stress-based approach. Thus, controlled blasting has not been used to predict liquefaction
triggering at a specific site. This may be less of a limitation, however, for strain-based and
energy-based approaches (e.g., Green, 2001; Green and Mitchell, 2004).
Vibroseis loading has been used to investigate pore-pressure generation characteristics in
situ (see, e.g., Chang et al., 2007; Cox et al., 2009), as well as the efficacy of various soil
improvement techniques (Chang et al., 2004; Stokoe et al., 2014a). A vibroseis truck applies
cyclic loading to the ground surface through a vibrating base plate (see Figure 5.9). Horizontal
excitation induces shear waves that can generate excess porewater pressures in the liquefiable
layers. The excess porewater pressures are measured by embedded pore-pressure transducers,
and the motions from the seismic waves are measured by accelerations or geophones. The
induced dynamic loading can be quantified by the cyclic shear strain amplitude, which is
computed from motions measured at two or more locations (Rathje et al., 2004), or by the cyclic
shear stress amplitude, which is computed from the derived shear strain and an estimate of the
strain-compatible shear modulus (Stokoe et al., 2014b). Unlike laboratory tests, neither the cyclic
shear strain nor the cyclic shear stress amplitudes induced by vibroseis loading are constant after
porewater pressures are generated. In this sense, vibroseis loading is more similar to the loading
applied during earthquake shaking than during uniform cycle laboratory tests. Testing is limited,
however, to relatively shallow depths (less than about 3 to 4 meters), and the loading level,
whether quantified in terms of strain or stress, is known only after the vibration measurements
are processed. Perhaps more importantly, the soil around the embedded instrumentation may not
remain undrained during loading because pore pressures are not generated at locations adjacent
to the instrumentation array, which results in hydraulic head differences that induce flow.
Depending on the hydraulic conductivity of the soil, this drainage effect may be significant,
although it has not been investigated thoroughly in vibroseis liquefaction testing. Despite these
limitations, vibroseis testing to directly measure the in situ liquefaction resistance has provided
valuable insights into liquefaction.
ALTERNATIVE APPROACHES LIQUEFACTION TRIGGERING ASSESSMENT 129

FIGURE 5.9 General test setup for an in situ liquefaction test using a vibroseis truck and embedded instrumentation.
SOURCE: Cox et al., 2009.

COMPUTATIONAL MECHANICS APPROACHES TO ASSESS LIQUEFACTION


TRIGGERING

Computational mechanics approaches to predict pore-pressure generation within an


effective-stress site response framework have been used to evaluate liquefaction triggering. Two
distinct approaches to predict earthquake-induced pore pressures have been employed (Beaty and
Perlea, 2011): (1) a loosely coupled approach that predicts pore pressures based on accumulated
strains; and (2) a fully coupled approach that uses a plasticity-based constitutive model to predict
both the stress-strain and the pore-pressure response of the soil.
Martin and colleagues (1975) were the first to develop the loosely coupled approach. This
model predicts plastic volumetric strain and then converts this value to an equivalent increment
in porewater pressure. Alternative forms for the relationship that predicts the plastic volumetric
strain have been proposed (see, e.g., Byrne, 1991). The required model parameters are developed
from drained laboratory tests measuring volumetric strain and are validated using results from
undrained laboratory tests measuring pore-pressure generation. Other models that can be used in
the loosely coupled approach relate pore-pressure generation directly with the induced shear
strain and number of cycles of loading (see, e.g., Dobry et al., 1982; Vucetic, 1986; and
Matasovic and Vucetic, 1993). Numerous investigators have developed energy-based models for
pore-pressure generation (see, e.g., Nemat-Nasser and Shokooh, 1979; Davis and Berrill, 1982;
Berrill and Davis, 1985; Liang, 1995; Figueroa et al., 1997; Green et al., 2000; Davis and Berrill,
2001; Polito et al., 2008). Models for the loosely coupled approach also have been developed
using the relationship between pore-pressure generation and the number of stress cycles of
loading (Seed et al., 1976; Wang and Kavazanjian, 1989), but these models are not very robust
130 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

for predicting liquefaction triggering because they require knowledge of the number of cycles
required to trigger liquefaction. The pore-pressure generation models discussed above, or slight
variants of them, have been incorporated within effective stress-based site response analysis
programs D-MOD2000 (Matasovic and Ordonez, 2011) and DEEPSOIL (Hashash et al., 2012).
These programs can be used to predict pore-pressure generation within a given soil profile for a
specified earthquake input motion.
The fully coupled approach uses a constitutive model developed within a plasticity
framework to predict the stress-strain and pore-pressure responses within the soil (Chapter 8).
Examples of constitutive models developed for this purpose include those developed by Prevost
and colleagues (1985), Iai and colleagues (1992, 2011), Cubrinovski and Ishihara (1998), and
Dafalias and Manzari (2004), as well as models named PDMY02 (Yang et al., 2003, 2008),
UBCSAND (Byrne et al., 2004), and PM4-Sand (Boulanger, 2010; Boulanger and Ziotopoulou,
2013, 2015). Some of these models have been implemented in open source computer programs
(e.g., OpenSees) or commercially available (e.g., FLAC) computer programs. These constitutive
models typically have a large number of constitutive parameters, but several researchers have
developed practical recommendations that relate the model parameters to typical information
available from a site characterization. The PM4-Sand model (Boulanger and Ziotopoulou, 2013),
for example, has been methodically and transparently calibrated to produce responses consistent
with field and laboratory observations of liquefiable soil behavior. The recommended model
parameters from such a process are selected so that the models are calibrated to provide
numerical results consistent with accepted empirical models for liquefied soil. An example of
this calibration process for UBCSAND is shown in Figure 5.10.
In general, all of the pore-pressure generation models discussed above have been
implemented within dynamic response computer programs, and their use requires that input
acceleration-time histories be specified. They can provide insight into the process of pore-
pressure generation within a soil deposit during seismic loading. Nevertheless, there are a couple
of unresolved issues when using these models to predict liquefaction triggering. For one thing,
there is little documented experience of interpreting liquefaction triggering analyses within the
context of an excess pore-pressure ratio (ru) because it is not clear whether a pore-pressure ratio
close, but not equal, to 1.0 represents liquefaction as inferred from field evidence. For another
thing, it is difficult to obtain the necessary soil properties and model parameters for the analyses.
Input of representative earthquake acceleration-time histories may introduce ambiguity into
a computational mechanics evaluation of liquefaction triggering. Different time histories may
lead to different conclusions regarding liquefaction triggering. Guidance on this issue is available
for structural engineers (e.g., on how many time histories to use and how to reconcile the results
from analyses using different time histories), but it may not be appropriate for liquefaction
assessment. ASCE/SEI 7-10 (ASCE, 2010) mandates that at least three time histories be used in
a time history analysis, and that when less than seven time histories are used, the maximum
response from among the individual analyses be considered. When seven or more time histories
are used, the average response from among the analyses can be used for design. But, given that
the pore-pressure ratio becomes asymptotic as it approaches 1.0, the use of an average pore-
pressure ratio from multiple time history analyses is likely not appropriate. As noted above,
uncertainty regarding which pore-pressure ratio should be considered indicative of liquefaction
triggering may further complicate the interpretation of multiple time history analyses. It is also
not clear which approach to selecting representative time histories is correct for a liquefaction
analysis when the design acceleration is based upon a probabilistic seismic hazard analysis.
ALTERNATIVE APPROACHES LIQUEFACTION TRIGGERING ASSESSMENT 131

Arguments that using a suite of time histories that conform to a uniform hazard spectrum (e.g.,
that such analyses consider multiple earthquakes simultaneously and that a conditional mean
spectrum should be used [Baker, 2011]), as is common when site response analyses are
conducted for structural design purposes, are also relevant to liquefaction analyses. Hashash and
colleagues (2015) describe the use of two conditional mean spectra, one for a nearby relative
small magnitude earthquake and one for a larger but more distant earthquake, to evaluate the
liquefaction, lateral spreading, and seismic settlement for a site where the deaggregated hazard
data indicated that the earthquake hazard bimodal. In the absence of any guidance on how to
address these issues when a computational mechanics approach is used to assess liquefaction
triggering and its consequences, the design engineer must use professional judgment.

(a)

(b)

FIGURE 5.10 Calibration of UBCSAND models against (a) empirical estimates of CRR-SPT blow count relationship
and (b) overburden stress adjustment factor K. SOURCE: Beaty and Byrne, 2011.
132 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

The most critical shortcoming for the loosely coupled models is the lack of shear-induced
dilatancy in them. Without dilatancy and the associated reduction in excess pore pressure, the
loosely coupled approach may shield upper layers from the earthquake ground motions when
deeper layers liquefy. This “base-isolation” effect may appear beneficial, but it may not be
realistic in the field due to lateral variability, shaking due to non-vertically propagating shear
waves or surface waves, and the transmission of shaking during instances of dilation-induced
stiffening (Anderson et al., 2011). It is important to stress that it is never appropriate to use any
effective stress based site response analysis to predict the earthquake-induced CSR for use in a
simplified stress-based liquefaction triggering evaluation (see Chapter 4) because the simplified
approach ignores the effects of pore-pressure generation.
Effective stress site response analyses are being used in practice with increasing frequency
at liquefiable sites (Matasovic and Hashash, 2012). While these analyses have been used
primarily to evaluate a site-specific acceleration response spectrum at potentially liquefiable
sites, their use leads inevitably to consideration of whether or not the soil at the site will liquefy.
As discussed above, however, questions about the interpretation of the results and concerns
about the adequacy of these models as the soil approaches liquefaction limit their applicability
for predicting liquefaction triggering. Further, although studies have been conducted to evaluate
the ability of nonlinear, total stress site response analysis to predict site response (see, e.g.,
Stewart and Kwok, 2008), few studies have evaluated the ability of nonlinear effective stress
models to predict pore-pressure generation, liquefaction triggering, and surface ground motions.
Of particular concern is the ability of these programs to predict the pore-pressure response at
relatively large pore-pressure ratios, where shear-induced dilatancy can influence the response.
Validation exercises performed to date have used centrifuge data (see, e.g., Byrne et al., 2004)
and instrumented field sites such as the Wildlife liquefaction array site from the 1987
Superstition Hills earthquake and Port Island from the 1995 Hyogo-ken Nanbu (Kobe, Japan)
earthquake (see, e.g., Bonilla et al., 2005; Ziotopoulou et al., 2012). With the availability of
additional centrifuge data through NEES, as well as additional field measurements of pore-
pressure generation, these validation exercises need to continue. The LEAP project (Manzari et
al., 2014) is one example of continuing efforts to calibrate these types of analyses. To
demonstrate that these analyses can discriminate between liquefaction and non-liquefaction,
these validation exercises need to focus not only on accurately predicting liquefaction and pore-
pressure ratios close to 1.0 but also on predicting smaller pore-pressure ratios and non-
liquefaction under small to moderate levels of shaking.
6
Residual Shear Strength of Liquefied Soil

Key Findings and Conclusions

 The post-liquefaction triggering residual shear strength of soil is an important parameter in the
evaluation of consequences of liquefaction.
 The residual shear strength of soil that has liquefied depends on many factors, including the soil’s
grain size distribution, density, pre-earthquake state of stress, potential for void ratio redistribution,
and the configuration between the soil and structures within it.
 Owing to the many factors influencing residual strength, and that usually only a single value is
computed from a back analysis of a failure, residual strength is often interpreted as a system
characteristic rather than a material property.
 Several residual shear strength correlations are available for use in evaluation of liquefaction
consequences. Fundamental differences among approaches used to establish correlations include
whether the residual strength is dependent on the pre-earthquake vertical effective stress and
whether a correction to the normalized standard penetration test blow count is needed.
 Use of several different correlations when evaluating the residual strength is warranted considering
the different approaches to characterizing the residual strength and the substantial uncertainty
associated with the resulting correlations.
 Uncertainty can be reduced by reevaluating available data, collecting additional case history data,
laboratory testing (for fine-grained liquefiable soils), and physical model testing.
 Back calculation of residual field strength is complicated by the lack of field observations of flow
failures and lateral spreading due to liquefaction triggering, particularly in coarse-grained (e.g.,
gravelly) or fine-grained (e.g., low-plasticity silt) soils. More study is needed to inform
recommendations for establishing the residual strength of liquefied soils.

Liquefaction triggering often reduces the shear strength of a soil so much as to induce
damaging and potentially catastrophic displacements (e.g., flow sliding) and bearing capacity
failures. The value assigned to the post-triggering large strain shear strength of the soil (hereafter
referred to as the residual shear strength) is often the deciding factor when determining if the
consequences of liquefaction triggering are acceptable or if remediation is necessary.

133
134 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Theoretically, any continuous seam of liquefiable material could induce lateral spreading or flow
sliding regardless of how thin it is. Consequently, the value assigned to the residual shear
strength can often result in differences of millions of dollars in the cost of a project.
The residual shear strength of a liquefied soil depends on many factors, including the grain
size and grain size distribution of the soil and the initial density (or void ratio) of the soil. The
stresses in the ground prior to the earthquake, and the presence of overlying confining layers, are
also postulated to affect the residual shear strength of a soil. The post-triggering strength may
also change as consequences develop due to effects such as particle rearrangement, excess pore
pressure and stress redistribution, and intermixing of soils during deformation. In view of the
conditions that exist in the ground following liquefaction triggering, it is not surprising that
researchers have had difficulty both in obtaining consistent, reproducible estimates of residual
shear strength from field cases and in reproducing field conditions in the laboratory.
While laboratory testing has been used to estimate the steady state shear strength of soils
(discussed in Chapter 2), such testing generally is viewed as unable to replicate the residual shear
strength of liquefiable soil in situ (NRC, 1985; Seed, 1987; Stark et al., 1999). The steady state
shear strength is a measure of the post-triggering shear strength of a soil under laboratory
conditions. Field conditions often do not satisfy the conditions associated with the steady state of
deformation, however, because the porewater pressure and void ratio redistribution that may
accompany large deformation in the field violate the constant volume condition associated with
the steady state. Furthermore, stress redistribution may change shear and normal stress levels in
the ground significantly during flow sliding, and the constant strain rate requirement of the
steady state strength is rarely, if ever, satisfied.
Because of the inability to replicate field conditions in the laboratory, estimates of residual
shear strength are usually based on back analyses of case histories of lateral spreading and flow
sliding. Correlations between residual shear strength and either standard penetration test (SPT) or
cone penetration test (CPT) resistance have been proposed by a number of researchers, including
Seed (1987), Seed and Harder (1990), Ishihara (1993), Olson and Stark (2002), Idriss and
Boulanger (2008), Kramer (2008), Gillette (2010), Robertson (2010), and Kramer and Wang
(2015). Because of the many factors that affect residual shear strength, back-calculated shear
strength is often said to be a system characteristic rather than an inherent material property.
Given this assumption, the residual shear strength can be expected to vary from one site to
another because of system features such as geometrical configuration, stratigraphy, pre-
earthquake stresses, and characteristics of structural interfaces. In addition, a number of flow
slide case histories involve soils that slide into bodies of water. Hydroplaning traps water
underneath the sliding material, so there is no shear stress on the bottom of the sliding soil. This
can be significant when shear strengths are back-calculated.

RESIDUAL SHEAR STRENGTH OF SANDY SOILS

Several researchers have conducted field and laboratory studies to establish relations
between the residual strength and parameters such as SPT or CPT resistance, effective stress
prior to the earthquake, and fines content. Table 6.1 summarizes the studies on which residual
shear strength relationships most commonly used in practice today are based. It should be noted
that while these studies involved case histories of flow sliding, the results have sometimes been
applied in practice to evaluate the potential for both flow sliding and lateral spreading. Olson and
RESIDUAL SHEAR STRENGTH OF LIQUEFIED SOIL 135

Johnson (2008) found that the level of mobilization of the shear strength they back-calculated
from lateral spreading case histories was consistent with that calculated from flow slides by
Olson and Stark (2002). This finding is not generally accepted, however, and most investigators
have restricted their relationships to flow slides. The applicability of residual shear strength to
the evaluation of lateral spreading is discussed in additional detail in Chapter 7.

TABLE 6.1 Studies of Residual Strength Estimates Based on Flow Slide Field Case Histories
Source Input Parameters Remarks
Seed et al. (1988) Corrected SPT values Extension of Seed (1987)
with uncertainty.
Seed and Harder (N1)60-CS.
(1990)

Olson and Stark (2002) Pre-earthquake effective Database expanded from


stress and (N1)60 with no Stark and Mesri (1992).
fines correction. Also CPT Found no effect of fines
correlation. content.

Idriss and Boulanger Pre-earthquake effective Based on Seed (1987),


(2007) stress and clean-sand Seed and Harder (1990),
corrected (N1)60-CS. and Olson and Stark (2002).

Robertson (2010) Pre-earthquake effective Based on critical state


stress and clean-sand concepts and limited field
equivalent CPT. cases.

Wang (2003), Kramer Pre-earthquake effective Based on critical state


(2008), Kramer and stress and clean-sand concepts.
Wang (2015) equivalent CPT or SPT.

Seed (1987) proposed a correlation between residual shear strength and SPT resistance that
related the shear strength back-calculated from 12 case histories of flow sliding to what he
termed the equivalent clean-sand normalized SPT blow count, (N1)60-cs, of the soil. The
equivalent clean-sand blow count is obtained using Equation 6.1, as follows:

(𝑁1 )60−𝑐𝑠 = (𝑁1 )60 + 𝑁𝑐𝑜𝑟𝑟 (6.1)

where the value of the correction factor, Ncorr, depends on the fines content of the sand, as shown
in Table 6.2.1 The value of (N1)60-cs used in Equation 6.1 should represent the pre-earthquake
condition of the soil, although in some cases the values used in establishing the correlation were
measured after the earthquake.

TABLE 6.2 Values of Ncorr Proposed by Seed (1987)


Ncorr 0 1 2 4 5
% fines 0 10 25 50 75

1
It is important to note that the correction proposed by Seed (1987) to obtain the equivalent clean-sand blow count
for the residual strength correlation is different from the clean-sand correction used for the triggering correlations
discussed in Chapter 4.
136 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Using additional case histories and reanalysis of some of the case histories used in the
earlier studies, Seed and others (1988) and Seed and Harder (1990) developed revised residual
shear strength correlations that reflect uncertainty in the back-calculated values of residual shear
strength. Figure 6.1 shows the relationship proposed by Seed and Harder (1990). The correlation
presented in Figure 6.1 is still used by some engineers in practice, despite the wide band of
uncertainty associated with the residual shear strength. Figures 6.2 through 6.6 represent the
evolution of thinking regarding the relationship between residual shear strength and penetration
resistance since development of the Seed and Harder (1990) relationship.

FIGURE 6.1 Relationship between residual shear strength and estimated pre-earthquake equivalent clean-sand
normalized SPT blow count by Seed and Harder (1990). Note the wide bounds on the possible residual strength
suggested by these investigators as well as the need to apply a fines correction to the penetration resistance.

Castro (1987) reasoned that because soil density increases with increasing effective stress
and the residual shear strength increases with increasing density, the residual shear strength in
the field should increase with increasing pre-earthquake effective stress. Castro (1987, 1991) and
Castro and Troncoso (1989) applied this concept to several tailings dams in South America.
Using similar logic, Stark and Mesri (1992) correlated the ratio of the back-calculated residual
shear strength, Sr, to the initial vertical effective stresses, σ′vo (or σ′vc), to the equivalent clean-
sand normalized SPT blow count, (N1)60-cs, for 20 case histories of liquefaction-induced failure.
Olson and Stark (2002) used an expanded database to correlate the residual shear strength ratio,
Sr/σ′vo, to the normalized SPT resistance of the soil, (N1)60, and to the normalized CPT tip
resistance, qc1. They concluded that the residual strength ratio was independent of fines content
and recommended that no adjustment be applied to the SPT blow count for fines content. The
residual shear strengths back-calculated by Olson and Stark (2002) were generally consistent
with those presented by Seed and Harder (1990), with the notable exceptions of the case histories
published by De Alba and colleagues (1988) and Marcuson and colleagues (1979), for which the
values obtained by Olson and Stark (2002) were significantly lower than those obtained by the
original investigators. The relationship between Sr/σ′v and (N1)60 developed by Olson and Stark
(2002) is shown in Figure 6.2. Olson and Stark (2002) provide a similar plot for Sr/σ′v versus the
RESIDUAL SHEAR STRENGTH OF LIQUEFIED SOIL 137

normalized CPT tip resistance, qc1, and recommend that this be used in lieu of the SPT plot if
CPT data are available (due to the greater reliability of CPT penetration resistance values).

FIGURE 6.2 Relationships by Olson and Stark (2002) between normalized SPT blow count and post-liquefaction
(residual) strength ratio. The solid lines reflect the uncertainty in residual strength about the best estimate
(the dashed line) from field case histories. Also shown are previous residual strength relationships from Stark and
Mesri (1992) and Davies and Campanella (1994). Note that Olson and Stark (2002) concluded that a fines
content correction was not required for the normalized and standardized SPT blow count, (N1)60. A similar
relationship was provided for the normalized CPT tip resistance, qc1. SOURCE: Olson, S.M., and T.D. Stark.
2002. Liquefaction strength ratio from liquefaction flow failure case histories. Canadian Geotechnical Journal
39:629–647. © 2008 Canadian Science Publishing. Produced with permission.

Idriss and Boulanger (2007) used data from Seed (1987), Seed and Harder (1990), and
Olson and Stark (2002) to develop a correlation between the residual strength ratio (S r/σ′vo) and
the clean-sand normalized SPT blow count, (N1)60-cs. The correlation by Idriss and
Boulanger (2007), presented in Figure 6.3, bifurcates into two branches at normalized
clean-sand blow
counts greater than about 8 blows/30 cm: one branch for conditions where void
redistribution effects are expected to be negligible and one for cases in which void
redistribution effects may be significant. In practice, the relationship where void redistribution
effects could be significant
is usually used given the difficulty in establishing that void redistribution effects are negligible.
For (N1)60 values of 10 blows/30 cm or less, the best estimate line of Olson and Stark
(2002) agrees reasonably well with the Idriss and Boulanger (2007) relationship. Idriss and
Boulanger
(2007) also provide a relationship between residual strength ratio and equivalent clean-sand
normalized CPT tip resistance, qc1Ncs, drawn to be consistent with their (N1)60cs relationship.
Based on critical state concepts and a limited number of field case histories,
Robertson (2010) proposed the correlation between strength ratio and normalized CPT tip
resistance shown in Figure 6.4. Consistent with critical state concepts, Robertson (2010) limits
flow potential to
soils having a normalized CPT tip resistance less than about 70 atmospheres (lower bound
estimate). This value of normalized CPT tip resistance corresponds to a state parameter () of
approximately –0.05, consistent with findings that granular soils with  > –0.05 are expected to
experience strain softening and shear strength loss in undrained shear (Jefferies and Been, 2006;
Shuttle and Cunning, 2007). The Robertson (2010) curve in Figure 6.4 falls between the best
138 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

estimate line and lower bound line shown by Olson and Stark (2002) for CPT normalized tip
resistance for clean-sand values less than 60 atmospheres (normalized tip resistance for clean-
sand values less than about 6 MPa) and then trends sharply upward, becoming asymptotic at
roughly the same penetration resistance as the Idriss and Boulanger (2007) curve for conditions
when void ratio redistribution effects are expected to be negligible.

FIGURE 6.3 Relationship between residual strength ratio and equivalent clean-sand normalized SPT blow count by
Idriss and Boulanger (2008) for σ′vo (or σ′vc)< 400 kPa. Note the bifurcation of the curve at a blow count of around 10
into curves representing cases where void ratio redistribution is negligible or could be significant. Note also that these
investigators apply the same fines correction employed by Seed (1987) and Seed and Harder (1990) to the
penetration resistance. SOURCE: Idriss, I.M., and R.W. Boulanger. 2008. Soil Liquefaction During Earthquakes.
Earthquake Engineering Research Institute MNO 12. Oakland, CA: Earthquake Engineering Research Institute. With
permission from the Earthquake Engineering Research Institute.

FIGURE 6.4 Post-liquefaction strength ratio correlation with normalized cone penetration tip resistance with Class A
and B case histories superimposed. The class rating relates to the quality of the field investigations of the case
histories. Note that, consistent with Seed and Harder (1990) and Idriss and Boulanger (2007), but contrary to Olsen
and Stark (2002) and Kramer and Wang (2015), Robertson (2010) does employ a fines correction to the penetration
resistance. SOURCE: Robertson, P.K. 2010. Evaluation of flow liquefaction and liquefaction strength using the cone
penetration test. Journal of Geotechnical and Geoenvironmental Engineering 136(6):842–853. With permission from
ASCE. This material may be downloaded for personal use only. Any other use requires prior permission of the
American Society of Civil Engineers. This material may be found at
[http://ascelibrary.org/doi/abs/10.1061/%28ASCE%29GT.1943-5606.0000286].
RESIDUAL SHEAR STRENGTH OF LIQUEFIED SOIL 139

The previously described procedures (see, e.g., Olson and Stark, 2002; Idriss and Boulanger,
2007; Robertson, 2010) assumed the ratio of the residual strength to the pre-earthquake vertical
effective stress linearly related to a resistance parameter (e.g., SPT blow count or CPT
penetration resistance) value. That assumption implies that lines representing the critical state
void ratio, isotropic consolidation, and the undrained steady state shear strength are parallel on
plots of void ratio versus log effective stress. As was shown in Figure 2.5, however, the steady
state line generally is somewhat steeper than the consolidation curve is, which, in turn, suggests
that the soil becomes more contractive (i.e., the state parameter increases) with increasing
effective stress. Wang (2003), Kramer (2008), and Kramer and Wang (2015) defined the ratio of
the residual strength to the pre-earthquake vertical effective stress as a nonlinear function of the
normalized penetration resistance and the magnitude of the pre-earthquake effective stress. Like
Olson and Stark (2002), Kramer (2008) and Kramer and Wang (2015) did not find a systematic
variation in the residual shear strength with fines content. Figure 6.5 shows the variation in the
median estimate of the post-liquefaction strength, Sr, with (N1)60 and with initial vertical
effective stress predicted by Kramer (2008). Gillette (2010) also presented a model in which the
residual strength ratio varied with effective stress. Figure 6.6 relates the post-liquefaction
strength to both the penetration resistance and the initial effective vertical stress Weber
(2015). Note that these curves, recommended by Weber (2006) for use in deterministic
analyses, represent the 33rd percentile values from a probabilistic analysis.
The concept of a nonlinear, stress-level dependent relationship between the residual strength
and penetration resistance seems logical based upon critical state concepts. Nevertheless, the
problems noted previously in determining residual strength from laboratory test data and the
steady state shear strength must be kept in mind. More field data supporting this concept are
needed.

FIGURE 6.5 Variation in median of post-liquefaction strength with normalized SPT blow count and initial vertical
effective stress by Kramer (2008). SOURCE: Figure adapted by author (or authors, as appropriate) from Kramer, S.L.
(2008). “Evaluation of liquefaction hazards in Washington state.” Olympia, WA : Washington State Department of
Transportation, Office of Research and Library Services. Produced with permission.
140 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 6.6 Post-liquefaction strength ratio (S1/σ′v,o) with equivalent clean-sand normalized SPT blow count and
initial effective vertical stress by Weber (2015).

RESIDUAL SHEAR STRENGTH OF GRAVELLY AND FINE-GRAINED SOILS

There are few documented case histories of flow failures (and lateral spreads) involving
liquefied gravelly or fine-grained soils. Thus, there is a limited basis for the selection of residual
shear strength values for stability analysis for those types of soils. The few available data for
gravelly soils appear to be consistent with the assumption that the resistance to liquefaction
triggering and the residual strength of gravelly soils are similar to those of sandy soils of the
same equivalent penetration resistance 2 (see, e.g., Harder, 1988; Andrus and Youd, 1989).
Whereas available laboratory data indicate that the triggering resistance of gravelly soils is
comparable to that of sands of similar relative density, those data also indicate that the post-
triggering deformation potential of gravelly soils in undrained cyclic shear laboratory testing is
significantly less than that of sands at the same relative density (Kokusho, 2007). This suggests
that it may be conservative to assume that the mobilized residual strength of gravels during
lateral spreading is similar to that of sands of the same equivalent penetration resistance. There
are two kinds of soil classified as gravel: one with gravel particles floating in a matrix of finer
soils (e.g., sand), and one where the gravel particles form the soil skeleton (i.e., soil classified as
sand with enough gravel to form a stable soil skeleton, including possibly gravelly
sands). Laboratory testing has typically addressed the second type of gravel, and not much is
known about the first type. It is not clear that the conclusion about the deformation potential of
gravel being less than that of sand at the same relative density applies to gravel embedded in a
sand matrix. Because studying the behavior of gravelly soils in the laboratory requires large-
scale tests, additional case histories may be required to resolve this issue.

2
Equivalent SPT blow count corrected for the effects of gravel particles or obtained from correlations with the
Becker penetration test blow count.
RESIDUAL SHEAR STRENGTH OF LIQUEFIED SOIL 141

Besides the finding that there is little field evidence of flow sliding in fine-grained soils
(e.g., in low-plasticity silts and clays),3 the only potentially relevant data on flow sliding in these
types of soils is the observation that flow slides have occurred in fine-grained mine tailings
deposits (see, e.g., Castro and Troncoso, 1989; Castro, 1991). The relevance of findings on the
behavior of fine-grained mine tailings to the behavior of natural fine-grained soils is uncertain,
however. Additional research on this, and on the potential for flow sliding and lateral spreading
in fine-grained soils in general, is needed.

DEALING WITH UNCERTAINTY

The limited number of available case histories and the substantial uncertainty associated
with the load and resistance parameters characterizing flow failure and lateral spreading case
histories result in considerable uncertainty in the correlations between residual shear strength and
penetration resistance. In addition, the many factors that control the residual strength of liquefied
soil masses in the field suggest that there will always be large scatter in any correlation between
residual shear strength and a simple measure of penetration resistance or a combination of
penetration resistance and initial effective stress. The committee notes that Figure 6.1 by Seed
and Harder (1990) and Figure 6.3 from Olson and Stark (2002) provide some information on the
uncertainty of their residual shear strength estimates. Providing such information to help guide
engineers who wish to use these correlations is good practice. The model by Kramer (2008) and
Kramer and Wang (2015) allows for quantitative estimation of uncertainty in residual strength
considering scatter in the case history data and uncertainties in SPT resistance and vertical
effective stress. Even without considering the uncertainties in SPT resistance and vertical
effective stress, the standard deviation of the logarithm of the residual strength ranges from about
0.36 to 0.74 (a large range).
The large uncertainty in the evaluation of residual shear strength is often accommodated by
use of a conservative estimate of strength (e.g., Seed et al., 2003; Idriss and Boulanger, 2008).
While this may serve to ensure safe performance in future earthquakes, it may also result in
unnecessary and uneconomical levels of remediation. While it is likely there will always be
considerable uncertainty in residual strength relationships due to the factors discussed above,
some sources of uncertainty may be reduced or eliminated through further study, among them
reanalysis of available case history data, collection of additional and more detailed case history
data, and laboratory and physical model testing. Examples of ways to reduce uncertainties
include

 Evaluating the residual shear strength ratio as a nonlinear function of overburden stress,
as suggested by Kramer (2008) and Kramer and Wang (2015), and employing critical
state soil mechanics principles to guide how the relationship should look.
 For gravelly soils: evaluating reported incidents of flow sliding and lateral spreading
from the 1964 Good Friday Alaska Earthquake (McCulloch and Bonilla, 1970),
supplemented by additional field investigation as necessary.

3
Clays with very high sensitivity: for example, quick clays (Mitchell and Soga, 2005), also susceptible to large flow
slides, are usually triggered by static loading and are outside the scope of the present study.
142 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

 For fine-grained liquefiable soils: laboratory and physical model testing to establish the
relationship between the post-earthquake residual strength; the residual strength
measured by monotonic loading of cyclic laboratory tests after the soil liquefies; and the
large deformation shear strength of soil monotonically loaded in the laboratory using
triaxial and simple shear tests and from field vane shear testing; as well as to investigate
the relevance of data on the residual strength of fine-grained mine tailings to natural fine-
grained soils.

Most importantly, as discussed in Chapter 3, collection of additional well-documented case


histories of lateral spreading and flow sliding in all types of soils and geologic environments is a
critical need.
In the future, it may be possible to address the issue of residual shear strength as a system
response through the use of the computational mechanics tools discussed in Chapter 8 of this
report, combined with accurate modeling of the key features that determine the response of a
system involving soil liquefaction. This capability, however, is well beyond the present state of
the practice. At present, given the difference in approaches used to characterize the undrained
residual strength and the large uncertainty associated with residual strength values, the use of
multiple correlations when assessing the undrained residual strength of a liquefiable soil is
warranted.
7
Empirical and Semi-Empirical Methods for
Evaluating Liquefaction Consequences

Key Findings and Conclusions

 Simple criteria and procedures are available for screening sites for severity and damage potential as
a result of liquefaction triggering, but their limitations are not often stated explicitly, and the
applicability of many of them has not been assessed rigorously.
 Empirical and semiempirical models are available to evaluate the impacts of liquefaction, including
the potential for flow sliding, lateral spreading displacement, loss of bearing capacity, increased
lateral earth pressures, and buoyancy effects. Rigorous validation of these models is lacking
generally and validation is impossible in some cases due to a lack of appropriate case history data.
 Approaches to modeling loads induced on structures by liquefied soils are available, but lack
validation.
 The ability to evaluate liquefaction consequence models is limited by the availability of appropriate
case histories and physical model test data. Instrumentation, monitoring, and documentation of sites
with observable liquefaction impacts could yield additional field data.

The consequences of liquefaction are described and illustrated in Chapters 1, 2, and 6 and
can include vertical and lateral ground displacement; slumping and failure of embankments; loss
of foundation support; increased lateral loads on and reduced lateral resistance of structures and
their foundations; buoyancy uplift of buried structures; and modification of free-field ground
motions. This chapter focuses on empirical and semiempirical methods for evaluating these
consequences of liquefaction.
The complexity of the phenomena that follow liquefaction triggering has prompted the
development of a variety of empirical and semiempirical procedures used by practicing
engineers. These include screening procedures that employ damage indices to determine if the
consequences of liquefaction are of engineering concern; procedures to evaluate the potential for
lateral spreading, flow sliding, and slope instability; and procedures to evaluate the performance
of deep and shallow foundations, earth retaining structures, subsurface utilities, and buried
structures. While most of these procedures are logically formulated, and some are supported by
laboratory and model testing, there is little field verification of these techniques (with the

143
144 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

exception of some screening procedures). The lack of field verification may be attributed to the
paucity of appropriately documented case histories of liquefaction consequences.

SCREENING PROCEDURES

Several procedures have been developed to screen sites for the potential for damage
associated with liquefaction. Three types of screening procedures are used in practice: (1) a
procedure that establishes a threshold ground motion below which liquefaction damage is
expected to be inconsequential; (2) a procedure to screen sites for damage potential based on the
appearance of surface manifestations of liquefaction; and (3) procedures that use results from
liquefaction triggering evaluations to compute a numerical index for the severity (extent) of
triggering, which is presumed to correlate with liquefaction damage potential.

Threshold Ground Motions

The use of screening criteria to eliminate the need for a specific class of engineering analyses
(i.e., exclusionary criteria) is common among federal, state, and local public works agencies.
Federal Highway Administration (FHWA) guidance documents (see, e.g., FHWA, 2006) and
American Association of State Highway and Transportation Officials (AASHTO) specifications
also use earthquake magnitude and ground motion threshold criteria to determine if a
liquefaction consequence assessment is needed at a bridge site. AASHTO (2014) specifications
state that a liquefaction consequence assessment is not required for bridges designed to a life
safety standard if the 5% damped design spectral acceleration at one second corrected for local
site conditions, SD1, is less than 0.15 g, or 0.3 g if the liquefiable soil is low-plasticity silt. In
addition, the AASHTO guide specifications do not require a liquefaction consequence
assessment if groundwater is not within 50 feet of the ground surface and either the normalized
standard penetration test (SPT) blow count, (N1)60, is greater than 30 blows/30 cm; the
normalized cone penetration test (CPT) tip resistance, qc1N, is greater than 180 atmospheres; or
the normalized shear wave velocity (VS1) is greater than approximately 213 m/s (700 ft/s).
FHWA (2006) screens out the need for a liquefaction consequences assessment if SD1 is less than
0.25 and the site corrected spectral acceleration at a period of 0.2 seconds, SDS, is less than 0.35
g, or the earthquake magnitude is less than 6, or the earthquake magnitude is between 6 and 6.4
and either the normalized SPT blow count, (N1)60, is greater than 20 or (N1)60 is between 15 and
20 and SDS is less than 0.35 g. Despite the rather detailed specifics of these criteria, the study
committee is not aware of any documentation of the development of these screening criteria, nor
of any evaluation of their efficacy (e.g., comparison to case history data). Furthermore, these
criteria are specific to bridges designed in accordance with AASHTO standards, including a life
safety performance criterion, and should not be applied to structures other than bridges without
some demonstration of their effectiveness.
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 145

Screening for Surface Manifestations

The influence of a non-liquefiable capping layer (of a thickness H1) on suppressing the
surficial manifestation of liquefaction in an underlying liquefiable layer (of a thickness H2) was
quantified by Ishihara (1985) based on field performance data. Ishihara related the minimum
value of H2 necessary for liquefaction to manifest at the ground surface to H1 (i.e., to the non-
liquefiable surface capping layer thickness). Figure 7.1 provides the H1-H2 boundary curves
developed by Ishihara (1985) using a combination of field performance data from the 1976
Tangshan, China, earthquake, the May 26, 1983, M 7.3 Nihonkai-Chubu (Sea of Japan)
earthquake, and professional judgment. The point at which one of those boundary curves
becomes vertical represents a capping layer thickness beyond which no surface manifestation of
liquefaction is expected to be observed regardless of the thickness of the underlying liquefiable
layer. Later studies (Youd and Garris, 1995) supported the limiting thicknesses of capping layers
shown in Figure 7.1 for surface manifestations other than lateral spreading and the transient
ground movements referred to by Youd (1984) as ground oscillations. Experience in the 2011
Christchurch, New Zealand, earthquake indicates that while Figure 7.1 may be applicable to two-
layer systems composed of a non-liquefiable capping layer overlying a liquefiable layer, it is
difficult to apply this concept to stratified profiles with multiple liquefiable layers interbedded
with non-liquefiable layers (van Ballegooy, 2012, 2014b). In addition, this method may not
identify cases where settlement of a liquefiable layer does not result in surface manifestations
such as differential settlement or cracking of the surficial soil layer. If settlement due to
liquefaction beneath a non-liquefiable capping layer is of engineering concern, the engineer may
wish to evaluate the potential for such settlement using methods discussed subsequently and to
consider the engineering implications of the calculated settlement.
The concept described in Figure 7.1, however, is being used in New Zealand as guidance for
mitigating liquefaction impacts to structures on shallow foundations by creating a uniformly non-
liquefiable crust beneath the structure (van Ballegooy et al., 2015). While this concept has
proven applicable to cases with a uniform non-liquefiable crust, the case history data used to
support this concept are still limited, particularly with respect to earthquake magnitude and
intensity; consequently, continued review of the applicability of Figure 7.1 using data collected
in future earthquakes is warranted.
146 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

(a) (b)

FIGURE 7.1 Conditions of subsurface soil stratification required for surface manifestation of liquefaction.
(a) Chart showing thickness of non-liquefying capping surface layer required for surface manifestation
based on data from the 1983 Nihonkai-Chubu earthquake and peak ground acceleration (PGA) of 0.2 g.
and (b) a generalized chart showing boundary thickness of surface layers for other PGA values.
SOURCE: Ishihara, 1985.

Liquefaction Severity Indices

Liquefaction severity (damage potential) indices are intended to provide a measure of the
severity of surface manifestations based upon the cumulative liquefaction response of the profile.
Various severity or damage potential indices have been proposed in the literature, among them:
the Liquefaction Potential Index (LPI; Iwasaki et al., 1978); the Ishihara-inspired Liquefaction
Potential Index (LPIISH; Maurer et al., 2015c); the one-dimensional volumetric reconsolidation
settlement (S1VD); and the liquefaction severity number (LSN; van Ballegooy et al., 2012). These
terms are defined and discussed in the following sections.

Liquefaction Potential Index (LPI)

The LPI, proposed by Iwasaki and colleagues (1978), provides a depth-weighted index of
the potential for triggering of liquefaction at a site. As shown in Equation 7.1, LPI is computed
as
20 m
LPI = ∫0 F ∙ w(z) dz (7.1)
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 147

F = 1 – (FS) for FS ≤ 1 and F = 0 for FS > 1 (FS is the factor of safety against liquefaction,
obtained from a simplified stress-based liquefaction evaluation procedure); w(z)
is a linear depth weighting function given by w(z) = 10 – 0.5z (z is depth in
meters below the ground surface) and w(z) = 0 for z > 20 m.

The resulting index is therefore dependent on the thickness of liquefiable layers in the uppermost
20 meters, the proximity of these layers to the ground surface, and the amount by which the FS
against liquefaction is less than 1.0. As opposed to the Ishihara (1985) procedure described
above for determining the influence of a non-liquefiable crust over a single liquefiable layer, the
LPI applies to a profile with multiple liquefiable layers. LPI can range from 0 (no layers with an
FS less than 1 in the uppermost 20 meters of soils) to a maximum of 100 (the FS against
liquefaction is zero for all layers in the uppermost 20 meters). Analysis of data from 45 sites that
liquefied during the 1964 Niigata, Japan, earthquake showed that severe liquefaction occurred at
sites where LPI > 15, and little liquefaction occurred where LPI < 5 (Iwasaki et al., 1978).
Liquefaction hazard maps have been developed using the LPI framework for seismic regions in
many regions of the world (see, e.g., Sonmez, 2003; Papathanassiou et al., 2005; Baise et al.,
2006; Holzer et al., 2006a,b; Lenz and Baise, 2007; Cramer, et al., 2008; Hayati and Andrus,
2008; Yalcin et al., 2008; Chung and Rogers, 2011; Dixit et al., 2012).
In more recent work by Maurer and colleagues (2015c), a power-law depth weighting
function (i.e., 1/z, where z is depth) has been employed instead of the linear function used in the
original LPI framework. Maurer and colleagues (2015c) also modified the LPI framework to
account for the limiting thickness of the non-liquefiable capping layer central to the H1-H2 chart
developed by Ishihara (see Figure 7.1b). The revised severity index that incorporated these
modifications, named by Maurer and colleagues (2015c) the “Ishihara Inspired Liquefaction
Potential Index,” or LPIISH, was found to reduce the number of false-positive predictions (i.e.,
cases where manifestations were predicted but not observed) in 60 case histories from six
earthquakes in the Taiwan, Turkey, New Zealand, and the United States. Specifically, of the 60
case histories analyzed, 31% of no-manifestation cases had LPI < 5, while 100% of no-
manifestation cases had LPIISH <5. The LPI and LPIISH performed equally well, however, in
predicting true positives (i.e., cases where manifestations were observed as predicted), with 94%
of such cases correctly identified with either index.
LPIISH is computed as

20 𝑚 25.56
𝐿𝑃𝐼𝐼𝑆𝐻 = ∫0 𝐹(𝐹𝑆) 𝑑𝑧 (7.2)
𝑧

where

1 − 𝐹𝑆 𝑖𝑓 𝐹𝑆 ≤ 1 ∩ 𝐻1 ∙ 𝑚(𝐹𝑆) ≤ 3
𝐹(𝐹𝑆) = {
0 𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒

and

5
𝑚(𝐹𝑆) = 𝑒𝑥𝑝 ( )−1
25.56(1 − 𝐹𝑆)
148 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

where H1 is defined the same as H1 in Figure 7.1, and z is the depth to the layer of interest in
meters below the ground surface. Although LPIISH shows improved predictive capabilities over
LPI, it nevertheless can still yield incorrect predictions of liquefaction consequences. As a result,
there is still a need for further development of liquefaction damage indices if they are to be
considered reliable.

One-Dimensional Volumetric Reconsolidation Settlement (S1VD) and the Liquefaction


Severity Number (LSN)

Similar to LPIISH, LSN uses a power law depth weighting factor (i.e., 1/z) to determine the
cumulative liquefaction response of a profile. It also includes contributions from all layers that
have an FS < 2 (as opposed to the use of only layers with FS < 1when calculating the LPI). LSN
is computed as Equation 7.3, below:
𝜀𝑣
𝐿𝑆𝑁 = 1000 ∫ 𝑑𝑧 Equation (7.3)
𝑧

where εv is the calculated post-earthquake volumetric strain at depth z (in meters) expressed in
decimal form. To compute volumetric strain, van Ballegooy and colleagues (2012) used the
method proposed by Ishihara and Yoshimine (1992), as implemented by Zhang and colleagues
(2002) with CPT data. In the method by Ishihara and Yoshimine (1992), discussed later in this
chapter, the post-liquefaction volumetric strains are calculated as a function of the FS for
liquefaction triggering.
Because the volumetric strains from the Zhang and colleagues (2002) relationship are self-
limiting, LSN reaches a limiting maximum value as the FS decreases. This limiting value is a
function of relative density of the soil, resulting in a maximum LSN for a given soil profile
regardless of the ground motion intensity (e.g., peak ground acceleration [PGA]).
Similar to LPI, LSN correlated fairly well to the severity of surficial surface manifestations
observed during the 2010-2011 Canterbury, New Zealand, earthquake sequence in areas where
liquefaction did not manifest in the form of lateral spreading and for profiles that did not have a
clay capping layer or multiple non-liquefiable layers interbedded with the liquefiable layers
(Tonkin and Taylor, 2013; Maurer et al., 2014b, 2015a; van Ballegooy et al., 2012, 2014a). The
cumulative volumetric strain of the liquefiable layers in the soil profile, S1VD, was not as accurate
as either LPI or LSN. Nevertheless, a comparison of the accuracy of LSN versus the accuracy of
LPIISH was inconclusive (van Ballegooy et al., 2015).

Using Liquefaction Severity Indices

Liquefaction severity indices are useful in engineering practice for prioritizing sites for more
detailed investigation and for regional hazard mapping, but they are not a substitute for more
detailed analysis, and they do have limitations. Use of any of the methods described above
assumes that the severity of surficial manifestations of liquefaction serves as a proxy for
liquefaction damage potential. While this is likely a valid assumption for shallow foundation
systems and surficial or near-surface infrastructure, it may not be valid for deep foundation
systems. Furthermore, it has been found that damage indices based on the cumulative
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 149

liquefaction response or volumetric strain potential of a soil profile are poor predictors of lateral
spreading or flow sliding due to thin but continuous seams of liquefiable soil.
Even for non-lateral spreading manifestations, the severity thresholds reported in the
literature for any of the aforementioned severity indices need to be used with caution. For
example, several studies have shown that the correlation between computed LPI values and the
observed severity of surface manifestation can depend on the procedure used to compute the FS
against liquefaction and the fines content of the liquefied soil (Lee et al., 2003; Toprak and
Holzer, 2003; Holzer et al., 2005; Juang et al., 2008; Papathanassiou, 2008; Maurer et al., 2014b,
2015a,b). In addition, because the LSN framework and threshold severity values were developed
based on damage in a single geologic environment (i.e., Christchurch, New Zealand), their
application elsewhere needs to be validated by comparison to field performance data from
different environments.
Going forward, engineers need to keep in mind that these methods could be useful for
preliminary assessments, but they are not suitable for detailed assessment of liquefaction
consequences. Developers of severity index frameworks need to state explicitly the limitations of
their methods and the significant sources of uncertainty associated with them. Continued
calibration and validation of these methods against case history data—leading to improved
analyses and ultimately to quantification of the uncertainty associated with them—is warranted.

FLOW SLIDING

Empirical and semiempirical procedures for analyzing flow slide potential rely typically on
static limit equilibrium analyses (i.e., limit equilibrium analyses in which no seismic coefficient
is employed) using the post-liquefaction shear strength of the soil in the liquefied zones and
shear strengths appropriate for post-earthquake conditions elsewhere in the ground. If using
those shear strengths results in an FS less than 1.0 it is usually assumed that flow sliding will
occur if liquefaction is triggered. The resulting deformations are usually described qualitatively
as large. If flow sliding is predicted, remediation often is assumed to be required. In some cases
(e.g., for slopes supporting or adjacent to critical facilities), to provide an appropriate margin of
safety, an FS greater than 1.0 may be considered a necessary threshold value. In other cases,
however, designing for an extreme level of earthquake shaking is assumed to account for all of
the uncertainty in the assessment of the potential for flow sliding; an FS of 1.0 is then considered
appropriate. This is implicitly the case in FHWA and AASHTO Load and Resistance Factor
Design (LRFD) recommendations that call for a resistance factor of 1.0 for geotechnical design
for the seismic design ultimate limit state (see, e.g., Kavazanjian et al., 2011). The committee is
not aware of any quantitative assessment of reliability that can distinguish between these two
approaches. Development of a reliability assessment framework that allows comparison of
different simplified approaches to evaluating flow slide potential, and research on the parameter
values required to implement such an assessment, are warranted and would provide a consistent
means to incorporate uncertainties into flow sliding analyses.
No documented cases exist of flow failures that resulted from liquefaction of soils with an
equivalent clean-sand normalized SPT blow count, (N1)60-cs, greater than about 15 blows/30 cm,
or with a CPT tip resistance, qc1-cs, greater than about 85 atm. This raises the question of whether
this absence of flow slide case histories implies that the post-liquefaction strength of soil
increases sharply beyond these (N1)60-cs and qc1-cs threshold values (see Chapter 6) or if
some
150 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

other mitigating factor is involved. While it may be possible that liquefaction flow failures in
such soils simply have not been documented, the frequency of observations of liquefaction
triggering does not decrease significantly until (N1)60-cs reaches approximately 20 blows/30 cm
and documented incidents of liquefaction triggering and liquefaction-induced lateral soil
displacements include cases for (N1)60-cs up to about 26 blows/30 cm. The absence of case
histories of flow failure for (N1)60cs values greater than 15 blows/30 cm and qc1-cs greater than
about 85 atm, combined with the existence of case histories of triggering and lateral
displacement for (N1)60-cs up to about 26 blows/30 cm, is strong circumstantial evidence that
liquefaction-induced flow failure is unlikely when (N1)60-cs exceeds 15 blows/30 cm or qc1-cs is
greater than about 85 atm. 1 Nonetheless, in the absence of any alternative, when the post-
liquefaction strength must be assessed for soils with (N1)60cs values greater than 15 blows/30 cm
and qc1-cs greater than about 85 atm, the available correlations may be extrapolated beyond this
point. There are no circumstances, however, when it is appropriate to use an extrapolated
strength that is greater than the drained strength of the soil. Idriss and Boulanger (2008; see
Figure 6.3) indicate with dashed lines the values extrapolated beyond (N1)60-cs equal to between
10 and 15 blows/30 cm. The curve for the case where void ratio redistribution effects are not
expected to be significant becomes asymptotic when (N1)60-cs is equal to about 17 blows/30 cm.
The curve for the case where void ratio effects are considered to be significant extends out to
values of (N1)60-cs equal to 30 blows/30 cm, even though there are no case histories for (N1)60-cs
greater than 15 blows/30 cm on the plot. Additional study on flow slide potential of soil
characterized by penetration resistance values corresponding to (N1)60-cs of greater than 15
blows/30 cm and qc1-cs values greater than 85 atm would allow greater confidence in the use of
extrapolated strength values and provide needed insight into the relationship between triggering
and flow of denser liquefiable soils, even though the uncertainty in the residual strength of soils
with high SPT resistance is high.

LATERAL SPREADING

If a site is susceptible to liquefaction triggering and assessment indicates that flow sliding is
not anticipated, the potential for lateral spreading must still be evaluated. The lateral spreading
case history databases include slope gradients as flat as 0.5%, so it is difficult to eliminate the
potential for lateral spreading based on slope gradient alone. Four methods to estimate lateral
spreading displacements are commonly used in practice (Idriss and Boulanger, 2008): (1) case-
history-based empirical correlations with field soil and site parameters; (2) integration of
permanent shear strains in the liquefied soil; (3) sliding block displacement analyses using limit
equilibrium analysis with residual shear strength in the liquefied zones; and (4) computational-
mechanics-based methods that employ two- or three-dimensional numerical simulations and
constitutive models for liquefying soil. The first three methods are discussed in this section.
Computational-mechanics-based approaches are described in Chapter 8. Probabilistic methods to
evaluate lateral displacement are also introduced in this section and discussed in more detail in
Chapter 9.
1
Most of the flow failure case histories are for slope gradients less than 10% and depths less than 10 meters. Caution
is warranted when applying this finding to steep slopes with large initial static shear stresses and for critical
structures such as large earth dams.
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 151

Case-History-Based Empirical Correlations

Empirical correlations by Rauch and Martin (2000), Bardet and colleagues (2002), and
Youd and colleagues (2002) are among the methods available for assessing the amount of lateral
spreading. Table 7.1 presents the key parameters used by these investigators. There are many
common parameters among the different methods, although the precise definitions of those
parameters and the collective set of parameters used by investigators vary. For instance, Youd
and colleagues (2002) and Bardet and colleagues (2002) used a similar set of parameters in their
respective lateral spreading models, chosen based on earlier work by Bartlett and Youd (1995).
Bardet and colleagues (2002) omitted the grain size parameters employed by Youd and
colleagues (2002), however, due to the difficulty in evaluating them from case history data.
Rauch and Martin (2000) developed a hierarchy of three models of increasing complexity. Their
regional model employed just four parameters: earthquake magnitude, distance from the site to
the fault, PGA, and the duration of strong ground motion. Their site model also included general
characteristics of the site (e.g., ground slope, height of the free face) in the regression equation,
while their geotechnical (Geo) model also included the depth to the top of the liquefied zone and
the depth to the liquefied zone with the lowest FS.
152 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

TABLE 7.1 Key Parameters Accounted for in Empirical Correlations for Lateral Spreading (indicated by
check)
Youd Bardet et Rauch and
et al. al. Martin
Parameter (2002) (2002) (2000)
a
Regional,
Moment magnitude   b c
Site, Geo
Regional,
Distance to fault  
Site, Geo
Depth to liquefiable
 Geo
layer
Ground slope   Site, Geo

Height of free face   Site, Geo


Distance from free
  Site, Geo
face
Regional,
Free field PGA
Site, Geo
Strong motion Regional,
duration Site, Geo
Thickness of
saturated granular
 
soil layers with (N1)60
≤ 15
Grain size 

Fines content 
a
Denotes a regional model that includes earthquake magnitude, distance from the site to the fault, peak ground
acceleration, and the duration of strong ground motion.
b
Denotes a site model that included the parameters above and general characteristics of the site (e.g., ground slope,
height of the free face) in the regression equation.
c
Denotes a geotechnical model that included the parameters above and the depth to the top of the liquefied zone,
and depth to the liquefied zone with the lowest factor of safety.

Figure 7.2 compares the observed and predicted displacements from the correlations by
Youd and colleagues (2002) and Bardet and colleagues (2002). Figure 7.3 provides observed and
predicted displacements from Rauch and Martin (2000). Figures 7.2 and 7.3 show that
predictions made using these models may be as much as two times greater and smaller than half
the measured displacement. In some of the models, the smaller the calculated displacement, the
greater is the discrepancy between observed and predicted displacements. These models provide
insight regarding the factors that influence lateral spreading potential and may be useful for gross
screening assessments of lateral spreading potential (e.g., liquefaction hazard mapping,
preliminary site assessments). Given the large uncertainty associated with the models, they are
not suitable for detailed engineering analysis. Nevertheless, quantifying the uncertainty
associated with these models to facilitate their use (e.g., to quantify the confidence limits
on anticipated lateral spreading displacements) would benefit the tchnical community.
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 153

(a)
(b)

FIGURE 7.2 Comparison of observed and predicted displacements of lateral spreads from SOURCE (a) Youd, T.L.,
C.M. Hansen, and S.F. Bartlett. 2002. Revised multilinear regression equations for prediction of lateral spread
displacement. Journal of Geotechnical and Geoenvironmental Engineering 128(12):1007–1017. With Permission
from ASCE. (2002) This material may be downloaded for personal use only. Any other use requires prior permission
of the American Society of Civil Engineers. This material may be found at
[http://ascelibrary.org/doi/abs/10.1061/%28ASCE%291090-0241%282002%29128%3A12%281007%29]. and (b)
Bardet, J.P., T. Tobita, N. Mace, and J. Hu. 2002. Regional modeling of liquefaction-induced ground deformation.
Earthquake Spectra 18(1):19–46. With permission from the Earthquake Engineering Research Institute.

Regional Model Site Model Geo Model

Observed Avg_Horz (m) Observed Avg_Horz (m) Observed Avg_Horz (m)


FIGURE 7.3 Observed and predicted displacements of lateral spreads from Rauch and Martin (2000) models.
SOURCE: Rauch F.A., and R.J. Martin II. 2000. EPOLLS model for predicting average displacements on lateral
spreads. Journal of Geotechnical and Geoenvironmental Engineering 126(4):360–371. With permission from ASCE.
This material may be downloaded for personal use only. Any other use requires prior permission of the American
Society of Civil Engineers. This material may be found at [http://ascelibrary.org/doi/abs/10.1061/%28ASCE%291090-
0241%282000%29126%3A4%28360%29].

Integration of Permanent Shear Strains

The permanent shear strain approach integrates the earthquake-induced permanent shear
strain potential over depth to calculate a lateral spreading displacement at the ground surface.
154 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Zhang and colleagues (2004) refer to this integrated lateral displacement as the lateral
displacement index (LDI). The LDI is calculated as Equation 7.4:
𝑧
𝐿𝐷𝐼 = ∫0 𝑚𝑎𝑥 𝛾𝑚𝑎𝑥 (𝑧) 𝑑𝑧 (7.4)

where zmax is the maximum depth at which liquefaction is expected, and max is the maximum
possible permanent shear strain (evaluated at each depth, z). The maximum possible permanent
shear strain at each depth, 𝛾𝑚𝑎𝑥 (𝑧), is related to the potential for liquefaction and the relative
density of the soil. The relationships developed by Ishihara and Yoshimine (1992) are commonly
used to estimate 𝛾𝑚𝑎𝑥 . These relationships, presenting 𝛾𝑚𝑎𝑥 as a function of the FS against
liquefaction and the relative density of the soil, were developed from laboratory testing data of
clean-sand specimens reconstituted at different relative densities. Larger values of 𝛾𝑚𝑎𝑥 occur for
smaller FSs and lower relative densities, but there is scatter in the data. The effect of fines
content on strain potential is typically captured in practice through the same fines adjustments
used for liquefaction triggering. The committee is not aware, however, of any demonstration that
this approach is valid.
Zhang and colleagues (2004) combined the relationships from Ishihara and Yoshimine
(1992) with the limiting shear strain values of Seed (1979) to calculate the LDI. Cetin and
colleagues (2009a) developed an alternative model to estimate 𝛾𝑚𝑎𝑥 using an expanded dataset
of laboratory test results that predicts 𝛾𝑚𝑎𝑥 as a function of blow count and the cyclic stress ratio
(CSR).
Comparison of computed LDI values with measured displacements at lateral spread sites
(see, e.g., Shamoto et al., 1998; Tokimatsu and Asaka, 1998; Zhang et al., 2004) have shown that
the LDI displacements have to be modified to account for site geometry factors such as the slope
or the distance to, and height of, a free face. For example, Zhang and colleagues (2004)
evaluated the ratio of the observed lateral displacement (LD) to the LDI and found that this ratio
increased with increasing ground slope (see Figure 7.4a) and decreased with increasing L/H,
where L is the distance from the free face and H is the height of the free face (see Figure 7.4b).
Even with corrections for ground slope and L/H, this approach does not appear to be more
precise than are models derived from field case histories (i.e., accuracy generally within a factor
of 2).
Faris and colleagues (2006) used an approach similar to the LDI to develop an equation for
the maximum horizontal displacement of a laterally spreading site as a function of the equivalent
clean-sand blow count, (N1)60-cs, the CSR, the slope gradient, , and moment magnitude, M.
These investigators first calculate a shear strain potential index (SPI) for each liquefiable layer as
a function of (N1)60-cs and CSR. The SPI represents the maximum shear strain anticipated in the
liquefiable layer and is based upon curves developed by Wu (2002) from laboratory test data
predominantly for clean sands. Similar to the calculation of the LDI, the SPI is multiplied by the
layer thickness and then summed over the site profile to obtain an estimate of Hmax. Based on
mean values from a Bayesian analysis of case history data, Faris and colleagues (2006) proposed
the following deterministic equation for Hmax:

𝐻𝑀𝑎𝑥 = 𝑒𝑥𝑝(1.0443 ln(DPImax ) + 0.0046 ln(α) + 0.0029 𝑴) (7.5)

where Hmax is in meters and  is shown as a percent.


EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 155

(a) (b)
FIGURE 7.4 Comparison between observed and predicted values of lateral displacement over the lateral
displacement index (LD/LDI) for sloping ground sites without a free face: (a) sloping ground case histories with SPT
data, and (b) free-face case histories with SPT data. SOURCE: Zhang, G., P.K. Robertson, and R.W.I.
Brachman. 2004. Estimating liquefaction-induced lateral displacements using the standard penetration test or
cone penetration test. Journal of Geotechnical and Geoenvironmental Engineering 130(8):861–871. With
permission from ASCE. This material may be downloaded for personal use only. Any other use requires
prior permission of the American Society of Civil Engineers. This material may be found
at [http://ascelibrary.org/doi/abs/10.1061/%28ASCE%291090-0241%282004%29130%3A8%28861%29].

Because this equation does not consider the possibility of a free face at the site, its
use appears to be limited to sites where there is no free face. These investigators also use a
fines content correction factor to adjust the blow count: that is, to evaluate (N1)60-cs, which is
unique to their method. The equation does illustrate the influence of magnitude and slope
gradient on lateral spreading, however, and shows that lateral spreading is possible even
when there is no slope gradient: namely, in level ground ( = 0). Based on comparisons
to the maximum displacement reported at lateral spreading sites, displacements predicted
using Equation 7.5 appear to have levels of accuracy similar to other shear strain potential
methods.
Given that permanent shear strain methods have accuracy of the same order of magnitude as
the empirical equations described previously, they are subject to similar limitations: that is, they
may be suitable for screening and preliminary analyses but are subject to considerable
uncertainty when used for detailed engineering assessments. Furthermore, these models cannot
capture the lateral displacement associated with a thin but continuous seam of liquefiable
soil. The limited thickness of a thin seam of liquefiable soil would result in calculation of very
small lateral displacements for even large permanent shear strain values, contrary to
observations of lateral spreading due to thin liquefiable seams in the field. An improved level
of accuracy and more information on the limitations and reliability of these models are
needed to facilitate the use of these methods for detailed engineering analysis. While
computational methods (discussed in the next section of this report) are not necessarily subject
to any less uncertainty than these empirical methods, they do allow for greater detail and
precision (though not necessarily greater
156 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

accuracy) in engineering estimates of performance and should be considered for important


projects as a source of additional information on the performance of engineered structures at a
site that may be subject to lateral spreading.

Sliding Block Analyses

In a sliding block analysis (see Box 7.1) for lateral spreading, the yield acceleration (ky) for
the potential sliding mass typically is established using pseudo-static limit equilibrium analysis
with the residual shear strength(s) (implying large, unidirectional strains) assigned to the
liquefiable layer(s) and post-peak large displacement strength(s) to non-liquefiable soil layer(s).
As noted in the earlier section on flow sliding, this type of analysis is appropriate only if the
static FS in a limit equilibrium analysis that uses these shear strengths is greater than 1.0. The
yield acceleration is then found as the seismic coefficient (i.e., the lateral force coefficient used
in the limit equilibrium analysis) that results in a limit equilibrium FS equal to 1.0. Predictive
relations for permanent displacement as a function of the yield acceleration and seismological
parameters (e.g., magnitude, site to source distance, PGA, peak ground velocity) derived by
regression against the results of numerous sliding block analyses are available in the literature
(see, e.g., Jibson, 2007; Saygili and Rathje, 2008; Bray and Travsarou, 2007; Anderson et al.,
2008; Lee et al., 2010; Rathje and Antonakos, 2011; Bardet and Liu, 2009; Urzúa and Christian,
2013; Lee and Green, 2015). In Newmark-type analyses, these predictive relations can be used
with ky evaluated—as discussed above—to assess lateral spreading potential in liquefiable
terrain. Alternatively, an actual Newmark analysis can be performed using one or more design
shear stress time histories. When actual time histories are employed, the time history is often
obtained as the shear stress time history at the depth of sliding from a site response analysis in
which potential modification of the ground motion due to pore pressure development is not
considered.
These analyses are sometimes assumed to yield conservative estimates of the seismic
displacement if ky is based on the assumption that the shear strength drops to the post-liquefied
or large displacement value immediately at the initiation of strong ground shaking (Kavazanjian
et al., 2011). Detailed comparisons, however, have not been made between observed lateral
spread displacements and those computed from sliding block analyses. Furthermore, the
displacements computed by a sliding block analysis are strongly influenced by the characteristics
of the acceleration-time history, and liquefaction is known to change the characteristics of the
acceleration-time history. While the accuracy of these types of analyses is unknown, they are
sometimes considered to be more appropriate for evaluating lateral spreading on discrete soil
layers than are other simplified methods (Olson and Johnson, 2008; Kavazanjian et al., 2011),
and the argument that they are conservative (because they assume that liquefaction occurs
immediately at the start of the earthquake) is logical.
Considering the alternatives for calculating lateral displacements due to liquefaction of a
thin seam (e.g., Youd and Bartlett-type analyses), Newmark-type sliding block analyses for
evaluating lateral spreading potential are likely to continue to be used for the near future.
Nevertheless, comparisons between the results of sliding block analyses and deformations
observed in documented case histories of lateral spreading, and perhaps from centrifuge
experiments, are needed to assess the accuracy of these types of analyses in order to enhance the
confidence associated with this type of analysis.
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 157

Probabilistic Design Acceleration

Many seismic hazard analyses today employ a probabilistically based horizontal PGA for
design. One challenge using a probabilistically based design acceleration is that there is no
unique earthquake magnitude associated with it. Similar to the method described in the Chapter 4
discussion of consideration of probabilistic design acceleration when calculating FS using the
simplified method (Idriss, 1985; Finn, 2007), the levels of lateral spreading associated with the
probabilistically based design acceleration can be evaluated using the deaggregated magnitude-
distance information by calculating a lateral displacement for each magnitude-distance pair,
multiplying that lateral displacement by the contribution of that magnitude-distance pair to the
design acceleration, and summing up those displacements. It should be noted, however, that this
is not a complete probabilistic representation of the lateral displacement because it ignores
contributions to the displacement from events with other recurrence rates. To truly evaluate
lateral displacement in a probabilistic manner, displacements must be integrated over the entire
hazard curve using the performance-based design methodology described in Chapter 9.

BOX 7.1
Sliding Block Analysis

Calculating earthquake-induced slope deformation using the analogy of a block on a plane was first
reported by Newmark (1965). In these analyses, often referred to as sliding block analyses, the sliding
mass (e.g., a non-liquefied soil layer) is modeled as a rigid block resting on a plane subjected to an
earthquake motion described by an accelerogram. The block and plane initially move together with no
relative displacement. When the inertia and gravity forces on the block exceed the sliding resistance
between the block and the plane, however, the block begins to slide in relation to the plane. The
acceleration at which the block begins to move relative to the plane is referred to as the yield acceleration
(ky).
When the block first begins to slide relative to the plane, the block acceleration is constant (and
equal to ky) and the plane is accelerating faster than the block. The sliding of the block does not stop
immediately when the acceleration of the plane drops below k y because the inertia of the moving block
must be dissipated by the contact shearing resistance. So, the block continues to move at a decreasing
velocity until the block’s velocity equals that of the plane, at which time the block and the plane once
again move together. This pattern of relative displacement, illustrated in Figure 1, is sometimes referred
to as a “stick-slip” motion. For any particular set of acceleration-time history and sliding resistance, the
result of the sliding block analysis is the cumulative displacement over the entire record. The basic
technique can be modified—at the price of increased analytical and computational complexity—to
account for both cohesion and frictional components of the sliding resistance; for inclined planes; for
vertical as well as horizontal acceleration; for loss of sliding resistance with increasing relative
displacement; and to eliminate the potential for “upslope” movement.
158 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 1 The Newmark sliding block concept for evaluating the earthquake-induced slope displacements. Darker
shaded areas in the plot of acceleration versus time represent the deceleration of the block after the ground
acceleration drops below the yield acceleration. Note that in a liquefaction-induced lateral spreading analysis, the
yield acceleration (ky) is evaluated in a limit equilibrium analysis using the post-liquefaction (residual) shear strength
of the liquefied soil. SOURCE: Kavazanjian et al., 2011.

LIQUEFACTION-INDUCED SETTLEMENT

Liquefaction-induced settlement can be attributed to a variety of mechanisms, including re-


sedimentation or reconsolidation (volumetric strain) of the liquefied soil; ground loss due to
venting of liquefied soil (i.e., sand boils or ejecta); settlement associated with lateral spreading
under zero volume change; and settlement due to soil-structure interaction ratcheting and bearing
capacity failure. Settlement associated with reconsolidation can be estimated quantitatively, but
no simple or simplified quantitative models are available to estimate the settlement associated
with ground loss due to venting of liquefied soil or to re-sedimentation of liquefied soil: namely,
settlement of liquefied soil where void ratio redistribution is significant beneath a capping low
permeability layer, as defined by Idriss and Boulanger (2008). The average settlement (i.e.,
vertical displacement) associated with lateral spreading movements can be evaluated using the
assumption of zero volume change of the sliding mass and the estimated lateral displacement.
More detailed evaluation of the vertical settlement that accompanies lateral spreading likely
requires using the computational-mechanics-based methods discussed in Chapter 8. Settlement
due to soil-structure interaction that results from racheting has been shown by Dashti and Bray
(2013) to be significant, but a method for predicting such settlements is not currently available.
Similarly, procedures to assess bearing capacity failure, discussed later in this chapter, do not
allow for assessment of associated settlement (although it is likely unnecessary to do so because
a predicted bearing failure will generally require remediation).
Empirical analyses of settlement due to post-liquefaction reconsolidation typically assume
free-field one-dimensional conditions. Under those conditions, the irrecoverable volumetric
strain associated with pore pressure dissipation is equal to the vertical strain. This vertical strain
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 159

can be integrated over depth to calculate the settlement due to reconsolidation. This calculation is
the same as that described in the earlier discussion on the use of S1VD as an index for damage
potential.
Various studies have shown that the post-liquefaction volumetric strain is influenced
predominantly by the maximum shear strain and the relative density of the soil (see, e.g.,
Tatsuoka et al., 1984; Tokimatsu and Seed, 1987: Ishihara and Yoshimine, 1992). Tokimatsu and
Seed (1987) developed a relationship between volumetric strain, SPT blow count, and CSR by
combining data relating shear strain to volumetric strain with previously developed relationships
between shear strain, SPT blow count, and CSR. Ishihara and Yoshimine (1992) related the
volumetric strain to the FS against liquefaction and relative density (see Figure 7.5). Zhang and
colleagues (2002) replaced the relative density term in Figure 7.5 with CPT tip resistance. These
relationships can be represented by equations that predict volumetric strain as a function of the
FS against liquefaction. Idriss and Boulanger (2008) also modified these relationships to be a
function of SPT blow count or CPT tip resistance.

FIGURE 7.5 Relationships to predict volumetric strain due to post-liquefaction consolidation as a function of the factor
of safety against liquefaction and relative density. SOURCE: Ishihara and Yoshimine, 1992.

Cetin and colleagues (2009a) developed a model for post-liquefaction volumetric strain as a
function of the clean-sand SPT blow count and the earthquake-induced CSR (see Figure 7.6a)
similar to the model developed by Tokimatsu and Seed (1987). Cetin and colleagues (2009b)
compared settlements computed by their model with measured settlements from 49 sites
encompassing several earthquakes (Figure 7.6b) and concluded that (1) the predictions were
generally within a factor of 2 of the observed values, and (2) their model provided the best
results among the several models they evaluated.
An important point is that the volumetric strain relationships discussed above are based
primarily on laboratory data from testing clean sands. In practice, engineers typically use the
equivalent clean-sand blow count or CPT tip resistance to calculate seismic settlement with these
models, assuming implicitly that the effect of fines content is accounted for in the fines
160 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

corrections for blow count or tip resistance. Studies of the effect of fines content on post-
liquefaction consolidation settlement are limited. Laboratory simple shear testing on relatively
undisturbed samples of sandy silt suggests that volumetric strains induced in liquefiable sandy
silt were approximately the same as volumetric strains estimated from equivalent clean-sand
blow counts (Reimer, 2007). These results are not necessarily contrary to findings from remedial
ground densification projects wherein settlement (and therefore densification) is induced by
liquefying the ground with techniques such as vibrating probes, weights dropped on the ground
surface, and detonation of buried explosives. According to Elias and colleagues (2006), the
effectiveness of these techniques in densifying the soil (i.e., inducing settlement) decreases with
increasing fines content. Few comparisons between post-densification settlements of clean sandy
soils and of silty soils in the field are available, however.

(a) (b)
FIGURE 7.6 (a) Cetin et al. (2009b) relationships predicting post-liquefaction volumetric strain as a function of the
clean sand SPT blow count and earthquake-induced CSR. (b) Comparison of observed and predicted post-
liquefaction settlement from Cetin et al. (2009a). SOURCE: (a) Cetin, K.O., H.T. Bilge, J. Wu, A.M. Kammerer, and
R.B. Seed. 2009b. Probabilistic models for cyclic straining of saturated clean sands. Journal of Geotechnical and
Geoenvironmental Engineering 135(3):371–386.With permission from ASCE. (b) Cetin, K.O., H.T. Bilge, J. Wu, A.M.
Kammerer, and R.B. Seed. 2009a. Probabilistic models for the assessment of cyclically induced reconsolidation
(volumetric) settlements. Journal of Geotechnical and Geoenvironmental Engineering 135(3):387–398. With
permission from ASCE. These materials may be downloaded for personal use only. Any other use requires prior
permission of the American Society of Civil Engineers. This material may be found at (a)
[http://ascelibrary.org/doi/abs/10.1061/%28ASCE%291090-0241%282009%29135%3A3%28371%29]; (b)
[http://ascelibrary.org/doi/abs/10.1061/%28ASCE%291090-0241%282009%29135%3A3%28387%29].
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 161

Observations show that in some cases damage from liquefaction-induced settlement may be
closely associated with sand boils, which are symptomatic of subsurface ground loss (i.e.,
venting of liquefied soil). Ground settlement associated with sand ejecta accounted for nearly
90% of the land damage in residential neighborhoods resulting from the 2011 earthquake in
Christchurch, New Zealand (van Ballegooy et al., 2014a). Differential subsidence associated
with ground loss due to sand ejecta in this earthquake is documented to have exceeded the
settlement expected due to free-field reconsolidation of the liquefied soil and to have been at
least partly responsible for differential settlement that damaged commercial structures in the
central business district in this earthquake (Bray et al., 2014). In addition to the structural distress
associated with sand ejecta, some areas in Christchurch impacted by large amounts of ejecta
subsided so that the land surface is now below high tide (Rogers et al., 2014). While there are no
quantitative analyses for total and differential settlement due to ground loss following
liquefaction, the liquefaction severity indices discussed previously appeared to correlate well
with observations of excessive settlement in Christchurch and may provide at least an indication
of how serious a problem liquefaction-induced settlement may be.

DEEP FOUNDATIONS

The potential consequences of liquefaction associated with deep foundations include loss of
vertical bearing capacity, loss of lateral stiffness and capacity, lateral loading due to lateral soil
displacements, and downdrag on piles due to post-liquefaction reconsolidation of soil. These
consequences may be divided into inertial effects (i.e., those that impact the dynamic response of
the pile foundation) and kinematic effects (i.e., those induced by soil displacement relative to the
foundation). Inertial and kinematic effects on deep foundations are often considered separately
because the available simplified methods cannot account for them as coupled phenomena. This is
justified because the peak inertial loads are likely to occur before the soil liquefies and begins to
flow because the intensity of the ground motion subsequent to liquefaction will be reduced and,
thus, the peak inertial load can be considered separately from the kinematic loads due to
liquefaction-induced displacements (Kavazanjian et al., 2011). The primary inertial interaction
effect prior to liquefaction-induced displacements is a reduction in lateral stiffness of the
foundation, which can be considered in the inertial loading analysis (as described subsequently).
The other effects cited above (loss of vertical and lateral capacity, lateral loading due to lateral
soil displacements, and downdrag on piles due to post-liquefaction reconsolidation of soil) will
occur only after triggering and, therefore, may be assumed to be independent of the dynamic
response of the foundation system. Alternatively, an advanced computation model such as those
discussed in Chapter 8 can be used to consider all inertial and kinematic loads together.

Vertical Capacity

The committee is not aware of formal procedures to account for the effect of liquefaction on
the vertical capacity of a deep foundation element (e.g., a pile or drilled shaft, referred to
generically herein as a pile). Seismic distress due to a reduction in vertical capacity induced by
soil liquefaction has never been documented. In practice, the vertical capacity of a deep
162 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

foundation is generally calculated as the sum of its side resistance and tip resistance. When
considered in an engineering analysis, the loss of vertical capacity of a pile in liquefied soil can
be accounted for by setting the side resistance of the pile in the liquefied zone to zero and by
reducing the strength of end bearing soils expected to liquefy to their post-liquefaction strength.
While there are few field or model test data to support this approach, in the absence of
information to the contrary, this approach is a reasonable means of accounting for the influence
of liquefaction upon vertical pile capacity based on the engineering mechanics of the problem.
Nonetheless, it would be prudent to validate this assumption through centrifuge model testing
and, should appropriate case history data become available, analysis of case histories.
There is a lack of consensus as to whether deep foundation elements in liquefiable soil are
subject to downdrag, wherein settlement of a non-liquefiable crust due to reconsolidation of
underlying liquefied soil pulls down on the sides of the pile. Logically, it is not necessary to
consider downdrag loads imposed by reconsolidation if the liquefied soil is above the neutral
plane of a driven pile. The pile has already been loaded by the soil above the neutral plane. If the
soil above the neutral plane liquefies and subsequently reconsolidates and reloads the pile, and
the soil below the neutral plane of the pile has not liquefied, then no significant settlement will
occur. If the soil below the neutral plane liquefies, as may be the case, for instance, for a bored
pile that has its neutral plane near its top, then settlement may occur due to the transfer of load
down to the deeper soils.
Establishing whether or not downdrag should be considered, and validation of the approach
generally used to consider it, is stymied by the lack of case history data on downdrag on piles in
liquefiable ground. The Juan Pablo II bridge performance during the 2010 Maule, Chile,
earthquake provides a rare case of seismic distress of a pier-supported structure due to reduction
in the vertical capacity induced by soil liquefaction (Ledezma et al., 2012). More case histories
of this effect are needed, however. Appropriate instrumentation and monitoring systems need to
be installed to capture field evidence of downdrag in areas where seismic loading resulting in
liquefaction is anticipated. Until appropriate data are available to evaluate whether downdrag is
of concern, it is prudent to consider downdrag when evaluating deep foundation performance.

Lateral Stiffness

The reduced lateral stiffness of deep foundations in liquefied soil is accounted for in a
simplified manner by reducing the lateral resistance of the liquefied soil in a p-y lateral load
analysis. A p-y analysis treats the pile as a beam on an elastic foundation with a load-dependent
or a displacement-dependent elastic response described by a curve relating the load, p, on a
segment of the pile to the corresponding lateral, displacement, y (Reese et al., 1974). Essentially,
each segment of the pile is supported by a spring, and the displacement-dependent spring
stiffness is characterized by the p-y curve. This type of analysis is akin to an equivalent linear
site response analysis in that iterations are performed until a displacement-compatible stiffness is
arrived at using the nonlinear p-y curve. Alternatives to this type of simplified analysis are
advanced computational methods (see Chapter 8) wherein either the soil is modeled as a series of
nonlinear springs attached to the foundation elements or a continuum soil model is used along
with structural elements and interface elements.
The reduced lateral resistance of liquefied soil in a p-y analysis is usually accounted for in
one of three ways: (1) by using a reduced undrained shear strength in the liquefied zone and
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 163

developing the p-y relationship by applying rules to develop p-y curves for the undrained
behavior of fine-grained soil (Abdoun, 1997; Wang and Reese, 1998); (2) by using a
multiplicative factor (i.e., a p-multiplier) that is less than 1 to modify the p-values of the p-y
curve for the non-liquefied soil (Ashford et al., 2011; Liu and Dobry, 1995; Wilson, 1998); or (3)
by using a constant pressure against the piles in the liquefied zone. Based on an analogy with
post-liquefactions stability assessments, it is often assumed that the residual shear strength can be
used as the reduced shear strength of liquefied soil. But, based on centrifuge test results, Abdoun
(1997) found that the lateral resistance of a single pile in liquefied loose sand could be
represented by assigning an undrained shear strength of approximately 1 kPa to the liquefied
soil, a value significantly less than what would be calculated using the post-liquefaction
undrained strength. The centrifuge tests were reinterpreted, and it was recommended that a
constant pressure of 10.3 kPa be applied to a pile in liquefied soil (Dobry et al., 2003). There was
good agreement reported between calculated and observed values when this assumption was
used to calculate pile bending moments with two field case histories given the assumption that
non-liquefied soil above the liquefied zone applied passive pressure to the foundation system
(e.g., pile and pile cap).
Data from centrifuge model tests on clean sand with a relative density of 40% frorm Liu and
Dobry (1995) was employed to develop the plot in Figure 7.7 of p-multiplier (called Cu in Figure
7.7) versus excess pore pressure ratio (ru).. Note that at an excess pore pressure ratio of 1.0,
representative of liquefaction, the p-multiplier is approximately 0.1, so the lateral resistance of
the liquefied soil is approximately 10% of its initial value. But, p-multipliers on the order of 0.25
have been reported for liquefied soil based on centrifuge model tests (Ashford et al., 2011).
Furthermore, assigning a reduced p-y resistance in the liquefied soil may be inappropriate
because of the generally gradual buildup of pore pressures during an earthquake, and because the
reduction in p-y resistance does not exceed 0.5 until the pore pressure ratio equals 70% (as
illustrated in Figure 7.7).
Centrifuge tests simulating lateral pile performance in liquefied denser sands (Wilson et al.,
2000) and lateral load tests on piles in soil liquefied by blasting (Rollins et al., 2005) show
stiffening due to shear strain dilation at large deflections as well as a different backbone curve
(i.e., for the locus of the tips of the hysteresis loops from uniform cyclic loading) than
conventionally assumed for p-y curves. Chang and Hutchinson (2013) have observed similar
behavior in shake table tests. A hybrid p-y model to account for this effect has been developed
by Franke and Rollins (2013).
There remain many uncertainties as to the nature of post-liquefaction p-y curves for laterally
loaded pile analyses. The influence of factors such as the gradual development of earthquake-
induced pore pressures, cyclic loading amplitudes, pile displacement induced pore pressure, pore
pressure dissipation, and the p-y resistance behavior of low-plasticity silts are among the major
uncertainties. Investigations of influence of these factors on the p-y behavior of liquefied soil
using centrifuge model and large-scale field tests and by study of field behavior of laterally
loaded piles in earthquakes captured through instrumented sites are needed to facilitate
development and validation of models for laterally loaded piles in different types of liquefiable
soil.
164 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 7.7 Reduction of lateral stiffness of a pile, Cu, as a function of excess pore pressure ratio, ru.
Note that at liquefaction, r u = 100% and Cu = 0.1 (lateral stiffness is 10% of the lateral stiffness in soil with
no excess pore pressure). SOURCE: Kavazanjian et al., 2011.

Loads Due to Lateral Spreading

Lateral spreading due to liquefaction can impart significant loads and displacements on deep
foundations as well as on the structures supported by the displaced foundation elements. This
type of loading is sometimes referred to as kinematic loading. In a structural analysis, kinematic
loading can be accounted for in a simplified manner by displacing the foundations of the
structure by an amount equal to the calculated free-field seismic displacements. This approach,
however, ignores the interaction between the displaced foundation elements and the structure
they support as the structure resists the foundation displacement. Properly accounting for such
soil-structure interaction effects is not amenable to the use of simplified models; it requires use
of computational models that include both structural and foundation elements (see Chapter 8).
As noted in the introduction to the section on deep foundations, in a simplified analysis,
kinematic loading of deep foundations due to lateral displacement of the soil surrounding a pile
subsequent to liquefaction is often decoupled from inertial loading from the superstructure (i.e.,
the two types of loading are considered independently) (POLA, 2010; POLB, 2012). Note that
this decoupled assumption applies only after the soil has liquefied, and that inertial and
kinematic interaction effects prior to the onset of liquefaction cannot be treated in a decoupled
manner. Two kinematic loading situations often encountered in practice are depicted in Figure
7.8. In the first case, the zone of liquefaction extends to the ground surface and the liquefied soil
tends to flow around the piles, applying relatively modest loads on the piles. In the second case,
the zone of liquefaction lies below a non-liquefied crust that exerts relatively large loads on the
piles. In both cases, the kinematic load imposed on the piles by laterally spreading soil can be
evaluated using a p-y analysis in which the soil stiffness in the liquefied zone is reduced in the
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 165

manner described previously. The kinematic load is applied to the pile by displacing the supports
for the p-y springs by an amount equal to the calculated free-field lateral displacement. This
approach to the kinematic loading problem is described by Martin and colleagues (2002) and
documented in an NCHRP Report (2002).
Ishihara and Cubrinovski (2004) used a reduced soil stiffness in the liquefied zone to
analyze the response of a pile-supported liquefied petroleum tank subject to lateral spreading in
the 1995 Hyogo-ken Nanbu (Kobe, Japan) earthquake. The situation analyzed by these
investigators is the case on the right-hand side of Figure 7.8, that of a non-liquefied crust
overlying the liquefied layer. In their analysis, Ishihara and Cubrinovski evaluated system
performance for stiffness reduction factors of 100 and 1,000. Good agreement with observed
behavior was reported for both reduction factors, leading the investigators to conclude that the
stiffness of the soil in the liquefied zone had little impact on the predicted structure response.

FIGURE 7.8 Idealized loading of deep foundations subjected to lateral spreading. SOURCE: USDOT FHWA.

Centrifuge tests to evaluate the p-y analysis approach to kinematic loading are described by
Brandenburg and colleagues (2007). McGann and colleagues (2012) describe a detailed
evaluation of this approach versus a three-dimensional finite element analysis. And McGann and
Arduino (2014) document a case history study using both a finite element and a p-y approach.
The use of the p-y approach for design is further described by Ashford and colleagues (2011).
Although this simplified approach to kinematic loading has been adopted in some seismic
design standards (POLA, 2010; POLB, 2015), many uncertainties remain in this type of analysis,
including, the validity of using the Newmark sliding block method to determine free-field
displacements; the uncertainties related to residual strength of liquefied soils and related p-y
curves; the uncertainties related to the thickness of the liquefied layer and the displacement
distribution in the layer; the influence of potential partial pore pressure increases in adjacent soil
layers; and the accuracy of the method in determining the distribution of pile bending moments.
Decoupling kinematic loading and inertial loading may not be appropriate for all situations,
especially for large-magnitude events where earthquake shaking is of such duration that lateral
spreading can induce kinematic loading on piles while strong shaking is still taking place. Some
166 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

design standards require coupling the two loading conditions (POLB, 2012; ASCE, 2014). Clear
guidelines on how to couple the kinematic and inertial loads are not provided in these design
standards, but Ashford and colleagues (2011) do provide guidance on their coupling.
Ultimately, data from instrumented deep foundations subject to lateral loading are needed to
validate these methods. It will take multiple earthquakes in extensively instrumented areas,
however, to provide those data. In the interim, some uncertainty can be reduced through
laboratory and field model testing supplemented by advanced numerical models (Chapter 8).
Sensitivity studies, such as those performed by Ishihara and Cubrinovski (2004), are needed to
assess system performance over the range of possible values for design parameters (e.g., the
reduced pile resistance in liquefied soil).

SHALLOW FOUNDATIONS

In most cases where liquefiable soils are present beneath or adjacent to the footprint of a
new structure, either the potential for liquefaction is mitigated or the deep foundations are used.
Shallow foundations are, in general, avoided in such cases, but before liquefaction analysis was
mandated by codes or was commonplace, structures underlain by liquefiable soil may have been
constructed. Furthermore, since shallow foundations are relatively inexpensive compared to deep
foundations, they may still be used when liquefaction-related deformations are expected to be
small (City of Los Angeles, 2014). Bearing capacity may be estimated using post-liquefaction
shear strength for the liquefied soil beneath the foundation. If a bearing failure is indicated, there
usually is no need to assess the associated displacements because remediation is generally
necessary.
The displacement of shallow foundations in liquefiable soil is often analyzed using the
decoupled kinematic loading approach for deep foundations discussed above. Liquefaction-
induced displacements, including lateral spreading and vertical settlement, are calculated and
applied to foundation soil as displacements (i.e., as a kinematic load), distorting the structure
without applying any inertial load. The forces and moments induced in the structure by this
distortion are then computed in a structural analysis. In a simplified analysis, the calculated
displacements are generally free-field displacements, ignoring any interaction between the
ground and the structure (e.g., any restraining effect the structure may have on the induced
displacements). Until recently, there were few data of any kind (e.g. model test data, field data)
with which to validate this approach. Computational and physical (centrifuge) modeling by
Dashti and Bray (2013), however, and case histories from the 2010-2011 Christchurch, New
Zealand earthquake sequence (Bray et al., 2014) have indicated that the decoupled approach can
significantly underestimate liquefaction-induced displacements. These data need to be used in
studies to develop appropriate simplified analyses for the performance of structures supported by
shallow foundations on liquefiable soils.

RETAINING STRUCTURE DAMAGE

Soil liquefaction behind an earth retaining structure will increase the lateral earth pressure
(i.e., the active thrust) on the structure. Liquefaction of soil in front of an earth retaining structure
will reduce the lateral resistance (i.e., the passive resistance) of the structure. Liquefaction of soil
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 167

beneath a retaining structure can result in loss of bearing capacity and lateral sliding resistance as
well as vertical settlement. These effects in lateral earth pressure have led to bulging, tilting, and
sliding of the retaining structures and damage to structures behind the retaining structures when
not accounted for in design. Damage to earth retaining structures in the Port of Kobe in the 1995
Hyogo-ken Nanbu earthquake, described briefly in Chapter 1, is a notable example of this kind
of damage.
Lateral earth pressures due to liquefied soil are generally accounted for using an equivalent
fluid pressure concept. Both active thrust and passive resistance are calculated as equivalent fluid
pressures based on the total unit weight of the liquefied soil (Ebeling and Morrison, 1993). This
approach has been found to be consistent with observations of damage to earth retaining
structures supporting liquefied soil in earthquakes (Iai, 1998), indicating the validity of this
approach.

UTILITIES AND BURIED STRUCTURES

Damage to utilities (e.g., pipelines, sewer lines and storm drains, communications cables)
due to liquefaction-related phenomena, including lateral spreading, settlement, buoyancy forces,
and ground oscillation (or transient ground movements without permanent ground deformation),
is common in large earthquakes in urban areas. Analyses of the impact of liquefaction on utilities
include screening analyses of system wide performance and analyses of the performance of
discrete components subject to liquefaction-induced ground deformations.
Screening procedures have been developed to evaluate the expected severity of damage to
utilities from liquefaction-induced effects for use in loss estimation and for prioritization of more
detailed system-wide assessments of damage potential. Loss estimation programs such as Hazus2
typically have such procedures built into them, and some public works agencies (e.g., the Los
Angeles Department of Water and Power) have their own in-house screening procedures. The
American Lifeline Alliance has developed procedures and equations to estimate the frequency of
repairs needed to pipelines as a result of earthquakes and permanent ground deformation (ALA,
2001). These take into account such factors as pipeline material, joint type, corrosivity of soils,
and the permanent ground deformation.
Such screening assessments are useful not only for loss estimation and prioritization of more
detailed analyses for system upgrades but also for estimating system outage times and in
planning for repair crew needs following a major earthquake. Every major earthquake provides
an opportunity to refine such analyses. Often, however, they are not suitable for site-specific
damage potential assessments.
Simplified analysis of the impact of liquefaction-induced ground deformations on pipelines
and other linear utilities is typically accomplished using the decoupled kinematic loading
approach discussed previously (i.e., by conducting a structural analysis in which calculated free-
field ground deformations are imposed on the structure without any inertial load). As noted
previously, this ignores the impact of the structure on the ground displacement—a generally
conservative assumption, as, due to its stiffness, the structure typically restrains the ground to
some extent.

2
See https://www.fema.gov/hazus.
168 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

A common observation in liquefied soil in earthquakes is buoyant uplift of underground


structures. Figure 7.9 shows a series of sewer manholes that buoyed up along a residential street
during the 2010 Darfield, New Zealand, earthquake (GEER, 2010). Buoyant uplift can be
accounted for by treating the liquefied soil as an equivalent fluid with a unit weight equal to the
total unit weight of the soil and calculating the uplift force based upon the volume displaced by
the underground structure (Tobita et al., 2012). This approach, however, is applicable only to
isolated, unrestrained structures that are not laterally restrained either by structural considerations
or by non-liquefied soil zones.
In the past decade, state-of-the-art research on the performance of pipelines—including
sewer, potable water, and fuel—in liquefiable ground has been conducted (see, e.g., O’Rourke,
2010; O’Rourke et al., 2014). Models for both screening and detailed analyses of the
performance of underground utilities in liquefiable soil were developed. Chou and colleagues
(2011) have illustrated the use of centrifuge modeling and advanced computational analyses to
predict the behavior of a sunken tube tunnel (the Bay Area Rapid Transit system tunnel) across
San Francisco bay. Chou and colleagues (2011) illustrated that buoyant uplift alone may not be
sufficient to predict the response of buried structures in liquefied soil if they are restrained
laterally. In such cases, more sophisticated soil-structure interaction analyses such as those
discussed in Chapter 8 are required to evaluate the performance of a buried structure in liquefied
soil. There have also been some salient observations of pipeline performance during earthquakes
in liquefied soil. A notable example is the observation that fusion-welded polyethylene pipe,
almost without exception, did not rupture when subject to liquefaction and lateral spreading in
the 2011 Christchurch, New Zealand, earthquake despite ground displacements of a meter or
more (O’Rourke et al., 2014).

FIGURE 7.9 Uplifted sewer manholes along Lower Styx Road, 2010 M 7.1 Darfield, New Zealand, earthquake.
SOURCE: From NSF-sponsored GEER Report #GEER-024. Dated 14/08/2010.
EMPERICAL AND SEMIEMPIRICAL METHODS FOR EVALUATING
LIQUEFACTION CONSEQUENCES 169

LIQUEFACTION-INDUCED MODIFICATION OF GROUND MOTIONS

The triggering of liquefaction and associated soil softening can significantly influence wave
propagation through the liquefied zone and, thus, also influence the nature of ground surface
motions. Until recently, few strong motion recordings were available at sites that liquefied, but
such data now allow empirical evaluation of the effects of liquefaction on ground surface
motions and validation of the ability of nonlinear, effective stress site response analyses to
predict this effect.
Youd and Carter (2005) examined acceleration response spectra 3 derived from ground
motion records at five instrumented sites in liquefiable soil. They found a general reduction of
short-period spectral accelerations (spectral period < 0.7 seconds) and amplified long-period
spectral accelerations (spectral period > 0.7 to 1.0 sec) at those sites compared to sites without
liquefaction. Hartvigsen (2007) investigated the effects of pore-pressure generation on ground
surface motions by comparing results from nonlinear, one-dimensional effective stress analyses,
and total stress analyses of hypothetical liquefaction sites, and found results in general agreement
with those from Youd and Carter (2005). Evaluation of field data by Gingery and colleagues
(2014) showed spectral acceleration amplification factors of about 1.8 to 3.1 in liquefiable soil
profiles, findings generally consistent with those of Youd and Carter (2005) and Hartvigsen
(2007). Gingery and colleagues (2014) also found modest amplification at spectral periods less
than 0.05 seconds, which they suggest may be attributed to “acceleration spikes that occur in
association with the dilational part of liquefaction phase transformation behavior”: that is, with
short-period spikes in the ground motion acceleration due to stiffening of the liquefied soil as a
result of its tendency to increase in volume when sheared. Anderson and colleagues (2011) make
a similar observation with respect to the impact of dilation in their effective stress analysis of
seismic site response (i.e., that it can amplify the short-period spectral content of ground
motions).
The amplification factors proposed by Gingery and colleagues (2014) provide a means of
assessing the potential impact of liquefaction on ground motions at a site in the long-period
range. These factors are based on a small set of data, however; therefore, they may not reflect the
full range of conditions encountered in practice. Effective stress site response analysis of the type
conducted by Anderson and colleagues (2011) may be able to capture short-period amplification
effects, though this type of analysis (i.e., an analysis that considers dilation during phase
transformation) is not widely available. The limited field data and analytical studies on the
effects of liquefaction on ground motion do not provide enough information for development of
a simplified empirical or semiempirical approach to predicting with confidence the effect of
liquefaction on site response, either before or after triggering. Similarly, nonlinear effective-
stress site-response analyses have not been well validated in the post-triggering range and,
therefore, need to be used with care. Additional ground motion recordings at borehole array sites
that record shaking below the liquefaction layers and at the ground surface are needed to validate
nonlinear effective-stress numerical models for site response.

3
An acceleration response spectrum is a plot of the maximum acceleration (referred to as the spectral acceleration)
of a linear single degree of freedom system versus the natural period of the system (referred to as the spectral
period). It provides the ground motion input to the most common type of structural analysis in earthquake
engineering, modal superposition, and may be considered as a unique signature, or fingerprint, of the earthquake
ground motion.
8
Use of Computational Mechanics to Predict
Liquefaction and Its Consequences

Key Findings and Conclusions

 Computational methods that use principles of mechanics and incorporate appropriate


constitutive relations offer a more detailed and flexible approach to predicting liquefaction and its
consequences than do empirical and semiempirical methods.
 Use of computational methods in engineering practice is hampered by uncertainties in their
accuracy and their ability to capture important phenomena, difficulties evaluating necessary
constitutive model parameters, and other complexities of their application.
 Rigorous and innovative experimentation and analyses of the behavior of liquefying soil, along
with laboratory and field validation of the results, could raise constitutive relations to a level
commensurate with computational abilities.
 Computational models need to be validated through comparison of calculations with previously
vetted, accurate, and precise physical models (e.g., centrifuge test results), well-documented
field case histories, or previously validated benchmark analyses that engage the important
features of problems under consideration.
 No current computational approach or computer code can address all aspects of the behavior of
liquefiable soil. Computer codes vary in their constitutive models for materials, special features
(e.g., the availability of structural elements), and idiosyncrasies during use.
 Research is needed on ways to implement advanced constitutive models and on incorporating
mixed formulations to solve simultaneously for deformations and pore pressures.
o Instability and flow problems characteristic of liquefied soil could be addressed with
meshless and discrete element methods used in conjunction with multiscale approaches
and depth averaged flow models.
o The smoothed particle hydrodynamics (SPH) method, the material point method (MPM), and
the discrete element method (DEM) hold promise for simulating liquefaction flow and its
consequences.

The study committee’s statement of task (see Box 1.4) directs the committee to explore the
sufficiency, quality, and uncertainties of various methods used to assess the potential for and
consequences of liquefaction triggering. Previous chapters describe the most common

170
COMPUTATIONAL MECHANICS TO PREDICT CONSEQUENCES 171

approaches applied in research and practice today. Those empirical and semiempirical
approaches, however, cannot account for all effects that arise from site-specific geologic details,
structural configurations, and ground motion characteristics. As a result, the committee explored
methods used more commonly for other engineering applications: for example, in the modeling
of landslides. Engineering mechanics offers a framework for fundamentally sound approaches to
modeling liquefaction and capturing its effects.
Computational methods that use the principles of mechanics and incorporate appropriate
constitutive relations1 applied to a properly characterized site provide a means to capture such
behaviors as pore-pressure generation, redistribution, and dissipation; void (porosity)
redistribution; soil and porewater flow; and soil-structure interaction effects. These methods, by
necessity, rely on constitutive relations that describe the response of a soil to an external load
(either force or displacement). When combined with appropriate generalizations of fluid flow
relative to the solid phase, computational models can be used to solve boundary-value problems
such as the deformation and pore-pressure response of a soil layer subjected to earthquake
loading. Whether for the purpose of establishing boundary conditions or validating modeling
results, site characterization and ground motion characterization conducted at a level of detail
appropriate to analysis performed through numerical modeling will affect the fidelity of
numerical modeling. Numerical models cannot be trustworthy representations of reality without
an understanding of the physical site and its ground motion response. The elements of site
characterization for case history assessments described in Chapter 3 are also appropriate to
inform numerical modeling.
Engineering-mechanics-based computational methods are accepted and used widely in
structural, mechanical, materials, and manufacturing engineering, in the modern seismological
prediction of earthquake ground motions, and in many geotechnical engineering applications not
involving liquefaction. Computational methods are also increasingly applied to liquefaction
analyses on projects where liquefaction-induced failure could threaten life safety or have
substantial economic or environmental impacts. However, the accuracy of those liquefaction
analyses is uncertain, important phenomena may not be captured by available models, and a
relatively high level of user awareness of the complexities, idiosyncrasies, and limitations of a
particular computational model is required for their proper application.
Computational mechanics models have been used to investigate consequences of
liquefaction in the small deformation or stable regime for a variety of geotechnical problems,
including the effect of liquefaction on ground motions (Xu et al., 2003; Peng et al., 2004; Lu et
al., 2004; Gyori et al., 2011; Kramer et al., 2011); lateral spreading (Elgamal and Yang, 2000;
Elgamal et el., 2002; Yang and Elgamal, 2002; Peng et al., 2004; Lu et al., 2004; Forcellini, et
al., 2013); settlement (Elgamal et al., 2005; Shahir and Pak, 2010; Gyori et al., 2011); and the
effect of liquefaction on structures such as foundations, lifelines, dams, and retaining structures
(Manzari, 1996; Elgamal et al., 2005; Marcuson et al., 2007; Shahir and Pak, 2010).
Use of computational models in cases where large deformations and potential instabilities
are associated with liquefaction is much less common and reliable than when deformations are
small and the overall system remains stable. This may be attributed to the fact that constitutive
relations often do not accurately capture the behavior of soil as it approaches liquefaction and
enters the post-triggering regime. Many available constitutive relations begin to deviate

1
A constitutive relation is a set of mathematical equations that describe the response of a specific material or element
to an external forcing load (or stress) or displacement (or strain).
172 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

substantially from some aspects of stress-strain behavior as soil approaches the liquefaction state
(i.e., at zero or very small effective stress), and they were never intended to be applied in the
post-triggering regime. While many engineering applications are not concerned with large
deformations and the post-triggering regime, new models capable of capturing post-triggering
features such as phase transitions (i.e., from solid-like to fluid-like behavior), fabric effects, and
rate dependence need to be developed and validated to provide a complete picture of liquefaction
effects.

COMPUTATIONAL LIQUEFACTION MODELING IN ENGINEERING PRACTICE

Geotechnical engineers now use a variety of empirical and approximate computational models
and codes to analyze liquefaction problems. The simplest are one-dimensional codes that rely on
effective stress formulations of soil stress-strain/pore-pressure generation under cyclic loading, and
on solving the one-dimensional wave equation to simulate nonlinear site response and the
associated pore-pressure evolution due to seismic loading. D-MOD2000 and Deepsoil (Matasovic
and Hashash 2012) are typical examples of software available to those in practice. Soil stiffness
and strength evolution due to pore-pressure generation and dissipation are accounted for via
backbone curves2 that are part of the constitutive model used in the code. The codes typically use
pore-pressure models that soften the soil monotonically. Because these models ignore dilation that
occurs often when the shear strain increases, they can over-soften the liquefied soil and lead to
excessive displacements and unrealistic “base isolation” effects: that is, soil layers above the
liquefied layer are isolated from the strong ground motions (Anderson et al., 2011). Some one-
dimensional codes—for instance, NOAH (Bonilla, 2001; Bonilla et al., 2005) and PSNL (S.
Kramer, unpublished data—account for phase transformation behavior of soils and they are,
therefore, more appropriate for capturing the behavior of soil that dilates after liquefaction
triggering. Those models are generally not available commercially, however. In addition, two-
dimensional codes such as FLACTM (Fast Lagrangian Analysis of Continua), 3 FLIP (Finite
Element Analysis of Liquefaction Program), 4 and Open System for Earthquake Engineering
Simulation (OpenSees;5 described later in this section) have been used successfully to run one-
dimensional analyses if the soils are constrained to behave as shear beams.
More advanced codes rely on two- and three-dimensional solutions to the coupled system of
equations that describe soil response in boundary-value problems. In engineering practice, one of
the most widely used two-dimensional programs for modeling soil liquefaction and its
consequences is FLACTM. FLACTM is a finite difference code that allows implementation of
user-defined constitutive models, includes structural elements and energy absorbing boundaries,
and can employ a Lagrangian large strain formulation to accommodate large displacements. One
user-developed constitutive model used widely to model sandy soil liquefaction and its
consequences in FLACTM is UBCSAND (Beaty and Byrne, 2011). Figure 8.1 shows a post-
liquefaction snapshot of FLACTM-simulated displacements in the Upper San Fernando Dam from

2
A backbone curve connects the tips of the hysteresis loops generated by uniform cyclic loading. In the absence of
cyclic degradation, it is analogous to the monotonic loading stress-strain curve for the soil.
3
See http://www.itascacg.com/software/flac.
4
See http://flip.or.jp/flip_english.html.
5
See http://opensees.berkeley.edu.
COMPUTATIONAL MECHANICS TO PREDICT CONSEQUENCES 173

the 1971 San Fernando earthquake. Note that the liquefaction-induced deformation of the Upper
San Fernando Dam, with a maximum value of approximately 3 meters, was significantly less
than that associated with the Lower San Fernando Dam (described in Chapter 1).
While the parameters of the UBCSAND constitutive model can be developed from laboratory
test data, Beaty and Byrne (2011) provide parameters that allow UBCSAND model results to be
consistent with the deterministic standard penetration test (SPT) and cone penetration test (CPT)
liquefaction triggering curves for clean sand (described in Chapter 4) for uniform cyclic loading.
Boulanger and Ziotoupolou (2012) implemented a modified form of a bounding surface constitutive
model (Manzari and Dafalias, 1997) called PM4Sand into FLACTM. Boulanger and Ziotoupolou
describe a methodical calibration process that allows their model parameters to be developed from
conventional soil property data, facilitating use of the model in engineering practice. Both
UBCSAND and PM4Sand were developed to model cyclic mobility and its effect on triggering and
post-triggering deformations; to the committee’s knowledge, their ability to capture post-triggering
flow deformations is still under debate, although PM4Sand has a critical state framework and would
seem to have no fundamental limitation against simulating flow deformation.
FLACTM, a finite difference approach, is sometimes considered on the opposite end of the
computational spectrum from finite element approaches such as that employed in PlaxisTM, 6
another commercial computer program widely used among engineering firms that do
computational-mechanics-based analyses. PlaxisTM has many of the same features as FLACTM
does, including structural elements, energy-absorbing boundaries, and user-defined constitutive
models (e.g., including UBCSAND). FLACTM uses both simple but dense finite difference meshes
and a solution scheme that treats all problems as dynamic problems. PlaxisTM uses more elaborate
elements (e.g., 15-noded triangles) to achieve a high degree of resolution. A significant difference
between these two programs is the method used to input an acceleration-time history to the
analysis. PlaxisTM uses the conventional approach of inputting an acceleration-time history to the
base of the soil layer. FLACTM requires input of a velocity-time history that represents only the
incoming seismic wave. Mejia and Dawson (2006) discuss a proper method of inputting an
earthquake time history to FLACTM. Discussion of their relative merits is beyond the scope of this
report, but these programs are well respected and widely used in the earthquake engineering
community. Matasovic and Hashash (2012) provide a brief general discussion of their use for site
response analysis considering pore-pressure generation as well as the need for calibration and
validation of such analyses.

FIGURE 8.1 Displaced shape and displacement vectors at the end of post-earthquake analysis of Upper San
Fernando Dam in the 1971 San Fernando Earthquake. This figure shows the ability of a common commercial
7
program to capture relatively large displacements. SOURCE: UBCSAND Constitutive Model on Itasca UDM website.

6
See http://www.plaxis.nl.
7
http://www.itasca-udm.com/media/download/UBCSAND/UBCSAND_UDM_Documentation.pdf.
174 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

OpenSees is an open source, finite element-based code for calculating the performance of
structures and geotechnical systems subjected to earthquakes. The code has an array of
sophisticated structural, geotechnical, and interface elements and models that allow it to analyze
a variety of soil-foundation-structure interaction (SFSI) problems, including those involving
liquefiable soils. A number of constitutive models capable of handling liquefiable soils,
including bounding surface (e.g., Manzari and Dafalias, 1997) and multi-yield surface family
models (Elgamal et al., 2003; Yang et al., 2003) are available under this numerical platform.
There is a steep learning curve associated with the use of OpenSees, however, and because it is
an open source program, limited technical support is available. Hence, OpenSees is used much
more commonly in research than in consulting engineering practice at this time.

ISSUES IN THE COMPUTATIONAL MODELING OF LIQUEFACTION PROBLEMS

Computational methods to analyze liquefaction problems solve the equations of motion and
deformation for assemblages of solid particles containing fully or partially fluid-filled pore
spaces. Mechanical principles require that stresses and displacements of the solid particle
framework, and fluid pressures and flows relative to those particles, satisfy the equations of
motion and of conservation of mass. Both the deformations of the solid particle framework and
the migration of porewater need to be accounted for, as does the transfer of momentum between
the solids and liquids. The nonlinear elasto-plastic nature of soil, the added complications of pore
fluid pressurization and flow, and the large deformations associated frequently with liquefaction
make problems in this domain among the most challenging in modern computational
engineering.
To simplify the analysis for liquefaction problems, it is generally assumed that the soil is
saturated, that the solid constituents (i.e., soil particles) are incompressible8 (although the soil
skeleton may be compressible), and that the motion of the interstitial fluid with respect to the soil
particle matrix is governed by diffusion equations. These assumptions are similar to the
conventional assumptions of Terzaghi consolidation theory, as extended by Biot (1941) to three-
dimensional deformation fields. Prevost (1980, 1985b), Prevost and colleagues (1985), Lacy and
Prevost (1987), and Zienkiewicz and colleagues (1990a,b, 1999) used continuum mixture theory
concepts to generalize Biot’s (1956a,b) dynamic poroelastic formulation into a form suitable for
advanced computational modeling.
In these approaches, only the equations of overall balance of linear momentum and of mass
conservation need to be resolved (Biot, 1941; Prevost, 1980; Zienkiewicz and Shiomi, 1984;
Andrade and Borja, 2007). The equations are discussed in Box 8.1. When combined, these sets
of equations furnish a coupled system within which the deformation of the solid matrix and the
pressure and transport of the pore fluid can be solved. Complexity here stems both from the
coupling between porewater pressures and solid matrix deformations and from the intrinsic
complexity of soil behavior. Setting the boundary and initial conditions is required for an
analytical or a numerical solution. Because of the complexity of the problem, analytical solutions

8
Some formulations use the bulk modulus of the fluid to approximate the pore-pressure evolution. But, in the fully
coupled analysis shown in Box 8.1, pore-pressure evolution is computed directly from the balance laws (Equations 1
and 2), without specifying the bulk modulus of the fluid.
COMPUTATIONAL MECHANICS TO PREDICT CONSEQUENCES 175

rarely are available. As a result, computational numerical solutions are typically employed in
practice.
In developing computational solutions for these problems, the analyst must make several
choices among alternative approaches, and he or she must consider the trade-offs between
analytical accuracy and computational efficiency. Some of these issues include how to discretize
the problem in space and time; whether to use small or large strain formulations; whether to
solve the coupled equations directly or to decouple them; whether to use a finite element or finite
difference model; how to deal with large deformations and discontinuities; how to address
questions of stability and flow sliding; how to model the interaction between the soil and any
pertinent structures; and, finally, how to validate the computational model. At this time, there is
no single computer code that addresses all of these issues satisfactorily, and there is no general
agreement within the geotechnical earthquake engineering community on how best to address
these issues. The following sections describe briefly the current state of the art and of practice.

BOX 8.1
Basic Balance Laws for Fully Saturated Soil-Fluid Analysis

Equations 1 and 2 are written using continuum mechanics conventions; a repeated subscript
indicates summation over the range of the subscript, and a comma indicates differentiation with respect to
the variable whose subscript follows the comma:

𝜎𝑖𝑗,𝑗 + 𝜌𝑔𝑖 = 𝜌𝑎𝑖 (1)


𝜌𝑓 𝑣𝑖,𝑖 = −𝑞𝑖,𝑖 (2)

In Equation 1, the balance of momentum involves the derivative (in space) of the total stress tensor (𝜎𝑖𝑗 ),
the density of the soil-fluid mixture (ρ), the acceleration due to gravity (𝑔𝑖 ), and the acceleration (𝑎𝑖 ) of the
soil skeleton (the formulation is more complicated in the mixture theory context, although some mass-
weighted acceleration 𝑎𝑖 can always be defined). In Equation 2, the balance of mass (flow) involves the
density of the pore fluid (𝜌𝑓 ), the derivative (in space) of the velocity (𝑣𝑖 ) of the solid skeleton, and the
derivative (in space) of the Darcy mass flux vector (𝑞𝑖 ). To close this system of equations, the effective
stress relation (plus constitutive equations linking the effective stress to the deformation of the solid
skeleton) and Darcy’s equation of flow are added. If the relative motion of the two phases is vigorous, a
Forchheimer formulation is possible (see, e.g., Mathias et al., 2008), with the pore-pressure gradient
balancing the sum of two terms, one being linear in the Darcy mass flux and the other quadratic in that
flux.

Discretization in Space and Time

The two classic approaches used to discretize a domain in space for a numerical solution are
finite differences (see, e.g., Moin, 2001) and finite elements (see, e.g., Hughes, 1987; Bathe,
1996; Zienkiewicz and Taylor, 2013). Accuracy of the solution in space increases with the
diminishing size of the “mesh” elements and with the complexity of interpolations within the
elements into which the domain is discretized. At the same time, however, computational
efficiency decreases. Thus, care must be taken to choose sufficiently small or complex elements
to provide an adequate level of accuracy while maintaining computational efficiency.
Time discretization (or integration) is typically accomplished using the Newmark family of
integrators, which can include, among others, the forward Euler scheme (explicit), backward
Euler scheme (implicit), and the Crank-Nicholson mixed scheme (Hughes, 1987), each with its
176 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

own advantages and disadvantages. Explicit schemes are typically used in practice because of
their simplicity, but care must be taken to choose sufficiently small time steps to provide
adequate accuracy and stability. Some nonlinear software (e.g., the well-known ABAQUS
system)9 includes provisions to address this limitation by varying the time step to maintain a
satisfactory energy balance at the end of each time increment. It is also important to note that
finer discretization will be required for similar reasons. Implicit schemes, while unconditionally
stable, are less efficient computationally and more difficult to implement in a computational
model. Hybrid schemes solve the fully coupled system of balance of momentum and flow using
an implicit-explicit predictor-corrector scheme. Given their unconditional stability, implicit
methods for time domain integration merit further development and increased use in
computational modeling of liquefaction. Crank-Nicholson schemes offer a compromise but can
produce spurious oscillations under certain conditions.

Small and Large Strain Formulations

Computational methods used to analyze liquefaction problems can be divided into methods
for small strains (typically less than ~1%) and larger strains, although there is no fixed
demarcation between what constitutes small and large strain. It must be remembered, however,
that large strain and large displacement are not synonymous. A body can undergo large
displacements while still experiencing small strains; for example, rigid body motion involves no
deformation at all. But, when the body deforms significantly, a large strain formulation may be
necessary. Small strain models are computationally simpler and more efficient than are large
strain models; therefore, they are generally used when the assumption of small strains is
adequate. Determining that adequacy, however, is somewhat subjective.
From a computational mechanics perspective, the distinction between large and small strain
methods is that small (i.e., infinitesimal) strain methods can be used when expressions that are
linear in the first derivatives of the displacements adequately describe the deformations and
rotations. In this case, the equations of equilibrium or of motion, which apply rigorously in the
deformed configuration, remain adequate when written without accounting for the geometry
change. Otherwise, large (i.e., finite) strain methods must be used. Most current formulations
presuppose small (i.e., infinitesimal) strains even when, in practice, they are used to simulate
finite strains.

Coupled and Uncoupled Solutions

A fully coupled solution scheme is one in which the changes in pore fluid pressures and
deformations in the soil particle framework are obtained simultaneously in each step of the
marching procedure. An uncoupled solution scheme is one in which, for example, the pore fluid
pressure is computed at some time and then allowed to dissipate without consideration of how
this affects the deformation and stress distribution in the solid constituents. Uncoupled solutions
for deformation and pore-pressure development tend to be less accurate but simpler to

9
See http://www.3ds.com/products-services/simulia/products/abaqus/latest-release.
COMPUTATIONAL MECHANICS TO PREDICT CONSEQUENCES 177

implement, and they are common in commercial computer programs. Uncoupled systems can be
combined into staggered systems in which, for example, the dissipation solution is carried
forward for one or more steps, the deformations and stresses are updated, the dissipation process
is carried forward for a few more steps, the deformations and stresses are updated, and so forth
and so on. Staggered solutions and fully coupled solutions are used less frequently than
uncoupled solutions, but they yield the most accurate results in computational analysis of
liquefaction problems. They require what is referred to as a mixed, or u-p (displacement–
porewater pressure) formulation in which both displacements and porewater pressures are
unknowns that must be computed.

Finite Element Versus Finite Difference Approaches

There is no perfect or ideal computational approach that can address all problems. Each
approach has its own strengths and weaknesses. The finite element method is very good for
modeling complicated geometries with different material properties in different regions. Indeed,
a motivation for developing the method was the difficulty encountered when trying to distort
finite difference meshes to represent the shapes of aircraft wings and fuselages (see, e.g., Gupta
and Meek, 1996). On the other hand, finite difference methods usually are more efficient at
propagating solutions through time. Both finite difference and finite element computer programs
are used in research and practice. Both methods can accommodate complex geometries,
boundary conditions, and finite deformations. Finite element-based computational methods are
by far the most common form of computational solutions in engineering mechanics.
Nevertheless, one of the most popular geotechnical programs used by practicing engineers
(FLACTM) is based on finite differences.
An important consideration in the development of finite element and finite difference
programs for seismic analysis is the treatment of boundary conditions to simulate infinite
domains. Simple fixed or free boundaries unrealistically reflect stress waves back into the
domain of interest and create what is known as a “bathtub effect.” Using large elements to
extend the geometry beyond the local domain of interest may also cause problems: the large
elements can have difficulty transmitting energy at wavelengths smaller than the element
dimension, thereby introducing distortions in the response at those smaller wavelengths. Care
should be taken when discretizing a problem with a mesh to ensure that frequencies below the
maximum frequency of interest can be propagated through those portions of the mesh that most
strongly affect the response. Material softening—as a result, for example, of pore-pressure
generation in liquefaction analysis—is also known to render computations sensitive to mesh size.
It has been shown, however, that this sensitivity can be ameliorated by the viscous effect
produced by fluid flow (Andrade and Borja, 2007). In selecting a finite element or finite
difference code for analysis of dynamic problems, the engineer should be sure that the code
selected handles the boundaries correctly.

Large Deformations and Discontinuities

When large deformations and discontinuities associated with liquefaction occur, both finite
element and finite difference formulations have to be enhanced to account for the deformations
178 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

and discontinuities in a manner consistent with the principles of continuum mechanics


(Belytschko et al., 2000; Holzapfel, 2000). Additionally, the spatial discretization techniques
(e.g., finite elements or finite difference mesh) of the computational model have to cope with the
large deformations resulting from the triggering of liquefaction. Computational modeling of
large deformations requires not only large deformation formulations for the governing equations
but also computational techniques to accommodate changes in geometry within the soil mass.
Examples of such techniques include Arbitrary Lagrangian Eulerian (ALE) techniques (Ponthot
and Belytschko, 1998); adaptive remeshing (Berger and Colella, 1989); or meshless methods
such as the material point method, smoothed particle hydrodynamics, and others (Sulsky et al.,
1994; Randles and Libersky, 1996). The finite elements themselves also have to be updated to
avoid the ensuing mesh distortions or discontinuities (Ponthot and Belytschko, 1998; Borja,
2000; Dolbow et al., 2001; Li et al., 2010). Additionally, the validity of constitutive relations—
developed typically to represent relatively small deformations—is questionable when the
deformations being modeled are large. More experiments (e.g., simple shear, triaxial, true
triaxial) are needed to characterize the response of liquefied soil to facilitate development of
models that account realistically for the behavior of soil after triggering of liquefaction. The
continuum approach has shown promising results in capturing some of the basic features
observed in the laboratory and physical models. Nevertheless, more laboratory and physical
models are needed to validate computational models and thereby increase the confidence in
simulation tools to address liquefaction problems accurately.

Stability and Flow Sliding

From a mechanics perspective, the unstable regime is one in which relatively small changes
in stress conditions yield a large response in strain (Hill, 1958; Nova, 1994; Bazant and Cedolin,
2010). In the liquefaction modeling community, the concept of physical stability is limited
generally to the “small” strain domain. Flow slides involve large strains and are considered to be
in the unstable regime. Other technical communities (e.g., those that study granular physics)
refer to stable regimes as “solid” stages and unstable regimes as “fluid” stages (Jaeger and Nagel,
1992). Modeling the transition from a solid to a fluid stage, or from a stable to an unstable stage,
has at least two basic requirements: (1) criteria to flag the onset of the transition, and (2)
computational schemes that can accommodate large strains and possible discontinuities in
displacement.
A circumstance of particular interest with respect to liquefaction consequences is lateral
spreading, where instability (i.e., flow failure) may not have occurred but where the soil has
undergone large displacements due to significant reduction in stiffness and strength (softening).
Lateral spreading in the field may also occur (Kramer, 1996; Elgamal et al., 2002) when the
effective stress approaches the zero-pressure region as a result of contractive material response
following multiple strain cycles. This results in significant softening but not necessarily in an
unstable situation.

Soil-Structure Interaction

Soil-structure interaction effects, including the impact of lateral spreading on pile


foundations and the lateral resistance of piles in liquefiable ground, are particularly important
COMPUTATIONAL MECHANICS TO PREDICT CONSEQUENCES 179

consequences of liquefaction (Chapter 7). Interpretation of aerial photographs following the 1964
Niigata, Japan, earthquake (Hamada and O’Rourke, 1992; O’Rourke and Jeon, 2000) indicates
lateral displacements of up to eight meters. While not necessarily indicative of flow-failure
instability, these displacements caused significant damage to structures. Computational modeling
of these interactions is complicated by the three-dimensional nature of the problems. While
several of the previously described computational platforms can accommodate three-dimensional
analysis, such analyses are labor and computationally intensive, and there are uncertainties
associated with three-dimensional generalizations of constitutive models. Therefore, many soil-
pile interaction analyses are conducted two-dimensionally, with the pile represented as a planar
element having an approximate lateral stiffness. A similar approach is applied in OpenSees,
which uses p-y curves for liquefiable soils (furnished by the use of “pysimple” elements), taking
into account pore-pressure generation and soil dilation and softening. While these analyses focus
on capturing the bending response of the pile, which is critical for most engineering applications,
it is sometimes necessary to represent the behavior of liquefied soil flowing around the pile.
Furthermore, both two- and three-dimensional analyses of soil-pile interaction require
consideration of the behavior of the soil-pile interface, another source of uncertainty. The
interface behavior between structural materials and both liquefied and non-liquefied soil is not
well understood. Generally, interface behavior is modeled using interface elements that employ
simplistic elastic–perfectly plastic constitutive models. For example, there is a need to develop
interface elements that can account for void ratio redistribution and the large deformations
resulting from the separation between the pile and the soil.
Improved computational efficiency for three-dimensional analyses, increased accuracy and
validation for three-dimensional constitutive models, and a better understanding of interface
behavior and the interface elements that model that behavior are required to obtain more accurate
representation of soil-structure interaction in liquefied soil.

Computational Model Validation

Fundamental to validating computational models is a comparison of calculations with


physical models (e.g., centrifuge test results), well-documented field case histories, or
benchmark analyses that have been previously vetted for accuracy and precision and that engage
the important features of the problem. For example, Peng and colleagues (2004) compare
centrifuge and finite element models of lateral spreading of a two-layer soil and its influence on
pile foundations. They successfully replicate the behavior of the soil and its interaction with the
pile under cycle loading. Boulanger and colleagues (2005) compare analysis of the restraining
effect provided by piles (referred to as “pile pinning effects”) in an embankment subject to
lateral spreading to centrifuge test results. This type of work exemplifies some of the challenges
related to predicting lateral spreading and its effect on structures: (a) the ability to capture
significant shear deformations as a result of material softening (1-10% shear strain); (b) the
ability of constitutive models to reproduce material behavior under numerous loading and
unloading cycles and as the effective stress approaches zero; and (c) the ability of the
computational scheme (e.g., finite elements) to cope with significant deformations, pore-pressure
generation, and even discontinuities in the displacement field in the soil itself or as a result of
separation between the soil and the pile. It is important to distinguish the need to capture realistic
material behavior and the resulting liquefaction effects (e.g., pore-pressure buildup) from
180 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

heterogeneities and complexities at the field level. While some continuum models have shown
promising results when compared with physical models, more efforts such as the Liquefaction
Experiment and Analysis Project (LEAP) (Manzari et al., 2014) and its predecessor, Verification
of Liquefaction Analysis by Centrifuge Studies (VELACS) (Arulanandan and Scott, 1993), are
still needed.
Currently, no computer code or analytical approach can address all aspects of the
deformation problems associated with liquefaction. Engineers working on projects for which
liquefaction effects can be expected to be important are well advised to use more than one
analytical or computational approach and to compare the results. Independent peer review of the
computational models and results should be incorporated into important projects.

CONSTITUTIVE MODELING OF LIQUEFIABLE SOIL

Solution procedures that contain the equations for motion and balance of mass for a coupled
system (see Box 8.1) must incorporate constitutive equations that relate effective stress changes
to strain increments. In addition, Darcy’s equation for fluid flow in porous media, or possibly an
extension of Darcy’s equation to account for deviations from pore-scale laminar flow conditions,
is used to complete the system of equations. Because of its central importance in computational
modeling, some remarks regarding constitutive modeling are appropriate.
Constitutive models need to simulate (a) undrained pore-pressure generation as a function of
the amplitude of shear loading and the material state (effective stress and density); (b) the effect
of initial shear stress (anisotropic initial loading); (c) transformation of the soil from a solid to a
liquefied state and its effects on stiffness; (d) the effects of various soil fabrics and fabric
degradation on stiffness; (e) stiffness and dilatancy evolution under cyclic loading; (f) the onset
of instabilities (diffuse and localized); and (g) the transition to the steady state (Kramer and
Elgamal, 2001). Capturing all of these features is beyond the capability of existing constitutive
models.
All constitutive models that attempt to describe the complexities of soil behavior need to
account for both elastic and inelastic deformations of the soil matrix. It is well documented that
the relations between stresses and strains in soils are nonlinear and inelastic and display
asymmetric response under compression versus extension (Schofield and Wroth, 1968; Muir
Wood, 1990). Perhaps the most widely used framework to account for this complex effective
stress behavior is elasto-plasticity, which accounts for both the recoverable (elastic) and the
irrecoverable (plastic) deformations of the soil matrix (Desai and Siriwardane, 1984; Prevost,
1985a; Lubliner, 1990; Simo and Hughes, 1998).
A number of available constitutive models represent the elasto-plastic stress-strain–pore-
pressure behavior of liquefiable soils (see, e.g. Prevost, 1985a,b; Iai et al, 1990; Towhata and
Ishihara, 1995; Parra, 1996; Manzari and Dafalias, 1997; Beaty and Byrne, 1998; Cubrinovski
and Ishihara, 1998; Kramer and Arduino, 1999; Elgamal et al., 2000; Li et al., 2000; Bonilla et
al., 2005; Ellison and Andrade, 2009; Beaty and Byrne, 2011; Boulanger and Ziotoupolou,
2012). These models typically focus on replicating the behavior of granular soils under cyclic
loading. Model features generally include state parameter (pressure and density) dependence,
dilatancy evolution, fabric degradation, inelasticity, anisotropy, steady-state behavior at large
shear strains, and tridimensional loading. By using an elasto-plastic constitutive model and the
mixed formulation for the coupling of deformation and porewater pressure described above, it
COMPUTATIONAL MECHANICS TO PREDICT CONSEQUENCES 181

ought to be possible to approximate the material and structural responses leading up to


liquefaction and the onset of large displacements observed in the laboratory and field (see Figure
8.2). A linear constitutive model is used sometimes for the porewater pressure.10 The bulk
modulus may be set equal to that of water for a saturated soil, or it may be set to a lower value if
the soil is unsaturated. The interaction between the effective stress and porewater fluid pressure
is governed by the coupled equations mentioned in Box 8.1.
Based mostly on observations of soil behavior in uniform cyclic laboratory tests,
constitutive models have evolved in their level of complexity. Some of the early constitutive
models (see, e.g., Martin et al., 1975) attempted to predict the evolution of the effective stress on
a cycle-by-cycle basis up to the triggering of liquefaction. (No attempt was made to predict soil
behavior after triggering.) Pore-pressure generation was predicted empirically rather than on the
basis of the coupling of volumetric deformation and pore pressure. Modern constitutive models
use coupled models based on effective stress-strain behavior and the associated pore-pressure
development in response to cyclic loads.

FIGURE 8.2 Finite element liquefaction modeling and prediction in the laboratory and field. (a) Undrained cyclic
triaxial test showing flow related to liquefaction for test conducted by Qadimi and Coop (2007) in loose sands; (b)
Simulations of undrained triaxial test using the Manzari and Dafalias (1997) constitutive model. Similar comparisons
in different stress paths have been reported in the literature (e.g., Andrade, et al., 2013); (c) field-scale simulation of
flow related to liquefaction for a bridge using realistic topography, seismic excitation, and finite elements, and an
appropriate constitutive response such as that shown in (b). Red colors imply larger cumulative vertical
displacements. SOURCES: (a) and (b) COURTESY: J. Andrade; (c) Arduino 2014 presentation to the committee.

10
Porewater pressure is related to volume change via an appropriate constant bulk modulus assigned to the pore
fluid.
182 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Most constitutive models based on elasto-plasticity use elastic response descriptors (e.g.,
elastic shear modulus, Poisson’s ratio), a yield surface, a plastic potential surface, a flow rule,
and a hardening rule. This framework requires a number of parameters to fully describe the
cyclic behavior of granular materials. It also relies on phenomenological observations in the
laboratory to formulate the hardening rules that control the internal (plastic) variables dominating
inelastic material behavior and to account for the basic mechanics such as dilatancy, anisotropy,
and fabric (Desai and Siriwardane, 1984; Muir Wood, 1990). Elasto-plastic models yield better
results when applied under the loading conditions for which they were calibrated (e.g., triaxial
compression/extension) than under other loading conditions. This is a direct consequence of
phenomenology (i.e., the constitutive equations are construed to represent observed behavior).
While the ability of elasto-plastic models to capture the complex behavior of soils is remarkable,
there is room to make these computational models more robust. Efforts are needed to develop
constitutive relationships that reduce phenomenology in favor of more fundamental
characterization of saturated granular flow. New developments in X-ray tomography and
micromechanics could shed light on these efforts. These models will more likely be adopted in
engineering practice when practical methods are developed to evaluate model parameters from
field test data (e.g., from SPT and CPT data) in a manner similar to that presented in Boulanger
and Ziotopoulou (2013).
Most constitutive models are based on the observed behavior of soil in laboratory tests at
relatively small strains (i.e., strains ≤ 10%); thus, they may be valid only up to liquefaction
triggering. To model post-triggering behavior of liquefiable soil accurately, constitutive models
that extend into the post-triggering large deformation regime are needed, but little is known
about post-triggering material response. Given the instability and large deformations that often
follow the onset of liquefaction, it is difficult to probe those areas of the stress space in the
laboratory. Effects such as phase transition, dilatancy, fabric evolution, pore redistribution, and
strain rate effects are important in capturing post-triggering behavior and are, therefore, features
that developers of models and numerical approaches have focused on to obtain more accurate
liquefaction and post-triggering predictions. More efforts are still needed in this direction,
especially at large deformations.

RECENT COMPUTATIONAL RESEARCH DEVELOPMENTS APPLICABLE TO


LIQUEFACTION ANALYSIS

Advanced computational techniques may be able to address such soil liquefaction issues as
void ratio redistribution and large displacement flow slides following the triggering of
liquefaction, and separations and discontinuities that sometimes accompany ground displacement
(e.g., separations between pile and soil and ground cracks in the upstream face of the Lower San
Fernando Dam; shown in Figure 8.3). Despite advances in computational capabilities, it is still
difficult to use computational techniques to capture pore-pressure generation in general loading
patterns beyond simple shear or triaxial conditions, to simulate large deformations, or to predict
post-triggering deformations consistent with residual strength and field observations. In
particular, when computational models attempt to capture the large deformations associated with
liquefaction triggering, they can misrepresent flow associated with liquefaction (and associated
displacements) and can fail to represent surface discontinuities associated with lateral spreading.
Meshless methods, such as the MPM (Sulsky et al., 1994; Mast, 2013; Ceccato et al., 2014) and
COMPUTATIONAL MECHANICS TO PREDICT CONSEQUENCES 183

FIGURE 8.3 (a) Mechanisms of failure for the Lower San Fernando Dam and (b) Finite element simulation of the
failure using mixed finite elements. SOURCE: Zienkiewicz, O.C., Y.M. Xie, B.A. Schrefler, A. Ledesma, and N.
Bicanic. 1990b. Static and dynamic behaviour of soils: a rational approach to quantitative solutions. II. Semi-saturated
problems. Proceedings of Royal Society of London A 429:311–321. By permission of the Royal Society.

the SPH method (Randles and Libersky, 1996), offer great promise for accommodating large
deformations through computational models based on finite deformation formulations. Other
alternatives include depth-averaged flow models (Iverson and George, 2014) and discrete
element methods (DEMs) (Cundall and Strack, 1979; Goren et al., 2010; O’Sullivan, 2011).
Meshless methods have the potential to capture complex material behaviors such as soil
skeleton–pore-fluid interactions—crucial in modeling liquefaction—and to represent the ensuing
large deformations. The continuum approach suffices to capture soil-fluid interaction using the
equations of balance of momentum and mass, but the continuum approach is not applicable
directly at the granular scale. At that scale, discrete element mechanics can be used to simulate
the behavior of the solid soil skeleton.
A computational mesh and nodes in the traditional sense are no longer necessary in
meshless methods, but implementing advanced elasto-plasticity models and incorporating mixed
formulations to solve the coupled dynamic and mass balance equations remain challenging. The
MPM and SPH procedures (both of which are meshless methods) still require continuum level
constitutive relations, and computational expense severely limits use of the DEMs. Some other
difficulties associated with the DEMs will diminish as computational power increases.
Despite their limitations, particle morphology and the coupling with fluid flow can be
captured more faithfully with discrete element methods. Also, meshless methods have much to
offer in capturing large deformations. Figure 8.4 shows simulations of flow problems using the
MPM and DEM platforms, displaying the capability of these methods to represent large
deformations resulting from flow associated with liquefaction or other phenomena. The
aforementioned methods provide alternative numerical solution architectures that may have
advantages over the traditional finite element and finite difference methods in modeling
liquefaction problems and are further described in this section.
184 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 8.4 Comparison of simulation of column flow problems using (a) the material point method (MPM) and (b)
the discrete element method (DEM). The granular material is initially in the box and then the box is removed
suddenly. Similar results have been published in the literature (e.g., Lim, et al.,2014). Both DEM and MPM
simulations shown are purely gravity driven, and in the case of the DEM they account for particle features such as
shape (discs versus angular) and friction. SOURCE: (a) Mast, 2013. Modeling landslide induced flow interactions with
structures using the material point method.; (b) COURTESY: J. Andrade.

Depth Averaging

The depth-averaged method reduces both the domain size and the discretization needs and,
therefore, increases computational efficiency in problems involving instability and flow. The
method simulates flow behavior by solving depth-averaged versions of the balance of linear
momentum and constant-volume equations. This technique, often used in the geosciences,
invokes the Mohr-Coulomb failure criterion to limit the shear stress at the base of the deforming
soil mass. Recent formulations also take into account the effective stress principle, observe
concepts of critical-state soil mechanics, and account for evolving solid volume fractions and
porewater pressures (Iverson and George, 2014).
Depth-averaged equations can be solved using conventional finite differences (control
volume), finite elements, or meshless techniques such as the MPM and the SPH method. The
depth-averaged method has been used to simulate large-scale debris flow problems under
controlled conditions and in natural settings (Iverson and George, 2014). Its two-dimensional
nature and efficient time integrators make this method capable of simulating flows at the
COMPUTATIONAL MECHANICS TO PREDICT CONSEQUENCES 185

kilometer scale with a good level of resolution in topography. It relies fundamentally on the
assumption that the depth of the flowing soil mass (the z-axis in Figure 8.5) is negligible
compared with its length, making it appealing for long runout flow. When the depth is a
significant fraction of the flow length, this approach has limited applicability.

FIGURE 8.5 Schematic of depth-averaged flow technique for simulating debris flow by solving the coupled system of
balance of linear momentum and mass equations. SOURCE: Iverson and George, 2014. Iverson, R.M., and D.L.
George. 2014. A depth-averaged debris-flow model that includes the effects of evolving dilatancy. I. Physical basis.
Proceedings of the Royal Society. With permission from the Royal Society.

Smoothed Particle Hydrodynamics

The SPH method (Randles and Libersky, 1996, and references therein) relies on
computational points over which a kernel is defined. A kernel is similar to a shape function in
finite elements in that it defines a family of hydrodynamic points that describe the pattern of
flow. The accuracy of the method depends crucially on the number of computational points. A
formulation of the SPH method can integrate the coupled system of balance of linear momentum
and mass equations to obtain simulations such as the one shown in Figure 8.6 of a landslide at
different time stations (Pastor et al., 2009). Fluid saturation and pore pressures are simulated, and
the formulation is more efficient than are other formulations such as those based on Eulerian
finite elements. The formulation is also depth integrated, making it less expensive
computationally.

Material Point and Discrete Element Methods

The MPM (Sulsky et al., 1994) and the DEM, illustrated in Figure 8.4, are promising
techniques to replicate material flow. Coupling these methods with pore-fluid flow to solve the
coupled system of equations for balance of momentum and mass is a challenge. Nevertheless,
these methods can be used to solve flow problems, such as the column drop test shown in Figure
8.4, with great accuracy. The column drop test simulates a granular material-filled container, the
walls of which are suddenly removed. The soil particles are allowed to flow as a result of
gravity. The boundary conditions for this problem impose large flow and deformations on the
soil mass that cannot be captured easily with mesh-based methods. The MPM and the DEM are
186 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 8.6 Simulated landslides for saturated frictional material considering pore-pressure dissipation. A dark-
colored mass of material seen at the top of the slope at time 8 sec fails and flows downhill until it reaches the bottom
of the slope by time 60 sec. Calculations are performed on a square mesh of about 1.5 km in each dimension.
SOURCE: Pastor, M., B. Haddad, G. Sorbino, S. Cuomo, and V. Drempetic. 2009. A depth-integrated, coupled SPH
model for flow-like landslides and related phenomena. International Journal for Numerical and Analytical Methods in
Geomechanics 33:143–172.

also capable of seamlessly simulating the interaction of soil with structures. Figure 8.7 shows an
MPM simulation of landslide interactions with a set of structures. Information about the forces
exerted on the structures by the flowing mass is among the computational results of this model.
Both the MPM and the DEM can also easily capture discontinuities in the deformation, including
complete separation of flowing masses into separate flows.

FIGURE 8.7 Evolution of landslide flow into protective barrier system using the MPM. Light (blue) color represents
landslide hitting a group of regularly arranged structures at time 6.25 sec. After 25 sec, the figure shows the
structures being completely engulfed by the landslide. SOURCE: Mast, 2013. Modeling landslide induced flow
interactions with structures using the material point method.

The DEM employs three main approaches to capture soil–pore-fluid interaction on a


multiscale basis. In order of increasing complexity, these approaches are (a) “dry” DEM
simulations imposing constant volume (undrained) constraints; (b) coarse-grid approaches in
which the fluid flow component is simulated using a discretization much larger than the average
particle size; and (c) fine-grid approaches in which the fluid flow component is simulated using a
discretization finer than the average particle size (O’Sullivan, 2011). The first approach is
straightforward and suffices for studying elemental tests,11 where undrained conditions hold,

11
Elemental tests are laboratory tests in which uniform stress conditions are imposed on a representative sample.
COMPUTATIONAL MECHANICS TO PREDICT CONSEQUENCES 187

roughly, locally (i.e., porewater redistribution effects can be ignored) and the soil and fluid
constituents can be assumed incompressible. On the other hand, fluid migration is relevant in
liquefaction modeling at scales larger than the element test (e.g., centrifuge, field scale). In these
situations, solving for fluid flow by explicitly using either coarse- or fine-grid approaches is
necessary to account for hydrostatic and drag forces, thereby accounting for particle-fluid
interactions.
Coarse-grid approaches tend to rely either on solving Darcy’s equation (Calvetti and Nova,
2004) or the averaged Navier-Stokes equation (Tsuji et al., 1993; Zeghal and El Shamy, 2004),
accompanied by the continuity equation. Specifically, Zeghal and El Shamy (2004) have used
the coarse-grid approach that solves the averaged Navier-Stokes equation to simulate the effect
of fluid flow on a cell of particles, thereby smearing the effect of the coupling over a large set of
particles. Figure 8.8 shows a two-dimensional schematic implementation of the averaged Navier-
Stokes equation (continuum analysis) coupled with the DEM (discrete analysis). This approach
generally works well in two dimensions, but its scale may be limited by its high computational
cost.
Fine-grid approaches rely on the solution of the full Navier-Stokes equation at the pore or
grain level using numerical techniques such as the SPH method (Potapov et al., 2001) or the
lattice-Boltzman method, where the Boltzman equation is solved over a fixed grid (lattice) that is
finer than the particle size (Cook et al., 2004; Sun et al., 2011a). The lattice-Boltzman approach
can be shown to capture the Navier-Stokes equation (Sukop and Thorne, 2006). Fine-grid
approaches offer higher fidelity, but they require significantly higher computational effort than
do the alternatives, making them best suited for highly detailed investigations rather than large-
scale simulations.
Further connection to the continuum scale emanating from DEM approaches necessitates
bridging techniques usually called multiscale approaches (Liu et al., 2006; Zhodi and Wriggers,
2005). DEM approaches rely typically on homogenizing or averaging (i.e., coarse-graining) the
interparticle contact forces using average effective stress tensor expressions such as the one
proposed by Christoffersen and colleagues (1981): see, for example, (e.g., Nicot et al., 2005;
Wellmann, et al., 2008; Andrade and Tu, 2009; Nitka et al., 2011; Guo and Zhao, 2014). On the
other hand, fluid flow, if resolved using fine-grid approaches, necessitates coarse-graining to
produce appropriate variables such as pore-fluid pressures (see, e.g., White et al., 2006; Sun et
al., 2011b). Multiscale methods connect these disparate representations of material behavior
across scales, from the grain scale to the field scale (Zhodi and Wriggers, 2005).
Multiscale approaches are finding increased use in geomechanics to inform continuum
models based on grain-scale resolution rather than relying on purely empirical relationships for
material property evolution (e.g., strength, permeability). Multiscale models are much less
computationally expensive than are full-resolution discrete mechanics approaches since they
resolve the fine scale at selected locations and at specific instances in time when such level of
resolution is needed (e.g., in areas of intense deformation). It is expected that with the increasing
level of computational power, use of multiscale approaches will increase in research and practice
to provide insight into the physics and mechanics of granular materials and, in particular, of
liquefaction phenomena.
188 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 8.8 Schematic of the average Navier-Stokes (continuum) coupled with DEM (discrete) approaches used to
simulate liquefaction behavior in granular soil deposits. SOURCE: Zeghal, M., and U. El Shamy. 2004. A continuum-
discrete hydromechanical analysis of granular deposit liquefaction. International Journal for Numerical and Analytical
Methods in Geomechanics 28(14):1361–1383.

Hybrid Approaches

It may be possible to develop hybrid techniques that use different constitutive models and
numerical techniques before and after the onset of liquefaction or large displacements. To
develop such methods, however, criteria indicating when the alternative constitutive models and
numerical techniques should be applied to cope with large deformations and different material
response after liquefaction triggering are needed (Nova, 1994; Borja, 2006; Andrade, 2009;
Buscarnera and Whittle, 2013). These criteria are typically based on material (internal) stability
analyses such as those proposed by Hill (1958).
Some hybrid approaches offer much promise and need to be developed further so their
fullest potential can be applied to liquefaction problems. In addition, a new generation of
computational models needs to be developed that can reproduce physical tests across scales
ranging from the laboratory to the field. This can be achieved by basing computational methods
on physics and mechanics to establish a more intimate coupling between computation, modeling,
and experimentation.
9
Performance-Based Evaluation and Design

Key Findings and Conclusions

 The uncertainties involved in the assessment of earthquake ground motions, system response,
physical damage, and loss make probabilistic methods for liquefaction consequence assessment
central to performance-based evaluation and design.
 Evaluation of liquefaction hazards and risks based on just one level of shaking, as is common in
current practice, provides an incomplete measure of actual hazards and risks.
 Probabilistic liquefaction hazard analyses in a performance-based design framework can account for
all levels of shaking, for uncertainties in the evaluation of the triggering of liquefaction, and for
uncertainties in the prediction of the consequences of liquefaction.
 Probabilistic liquefaction hazard analyses can directly predict the return periods of the consequence
levels themselves rather than a consequence level associated with a prescribed level of ground
shaking.
 To foster the adoption of performance-based liquefaction hazard evaluation in practice, computer
programs that perform the voluminous calculations involved in probabilistic performance-based
frameworks are required.

Performance-based design is a process in which the performance of a facility being designed


is evaluated over the entire range of possible loadings rather than for one or more discrete ground
motion return periods or events. The approach is intended to allow rational decisions regarding
appropriate design levels, including decisions related to the added value of increasing the seismic
resistance. Financial, construction, and operational issues are considered over the life cycle of the
facility. Seismic performance-based evaluation and design usually involves the probabilistic
descriptions of ground motions, system response, physical damage, and loss, including
casualties, direct economic costs, and indirect economic costs from loss of functionality (see
Figure 9.1). Performance-based evaluation and design have emerged within the earthquake
engineering community in recent decades as a preferred means to quantify the expected
performance of engineered facilities given the uncertain future demands placed on them by
earthquakes (NEHRP, 2006; PEER, 2010). Performance-based seismic design methods require
that design objectives and performance levels be quantified and that performance be predicted

189
190 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

over the full range of possible seismic events. Life cycle impacts and benefits of improved
performance can also be evaluated. This approach allows rational trade-offs to be made between
the construction costs required to achieve improved performance and the reduced life cycle
impacts associated with improved performance (e.g., reduced repair costs).

FIGURE 9.1 Schematic illustration of the components incorporated in performance-based evaluation and design.
SOURCE: Modified from Kramer, S.L., 2013.

The application of performance-based engineering and design and probabilistic evaluation


of risk and uncertainty has evolved significantly in recent years. These concepts are now central
to the management of dams and associated facilities by the U.S. Bureau of Reclamation and the
U.S. Army Corps of Engineers as well as numerous public and private agencies in Canada,
Europe, and elsewhere. Hartford and Baecher (2004) present a comprehensive review of the
background, theory, and application of performance-based engineering for the safety of dams.
The International Navigation Association (2001) guidelines incorporate performance-based
design concepts for port structures. The following pages describe the application of these
concepts in the specific context of seismically induced liquefaction for generic facilities.
As discussed in previous chapters, liquefaction hazards are generally assessed by evaluating
(a) the susceptibility of a soil deposit to liquefaction; (b) the potential for triggering of
liquefaction under the anticipated seismic loading; and (c) the consequences of liquefaction
triggering. Performance-based evaluation and design provides a means to translate consequences,
expressed in traditional engineering terms, into consequences expressed related to losses such as
casualties, direct and indirect economic losses, and loss of utility or functionality. Furthermore,
performance-based evaluation and design provides a means to consider the entire spectrum of
possible seismic events rather than probabilistically or deterministically posed discrete design
events.
Consideration of uncertainty is central to performance-based evaluations. In current
practice, liquefaction susceptibility, triggering, and consequences usually are deterministically
evaluated without explicit consideration of the uncertainty of the various components of the
analysis or the quantification of the uncertainty of the performance metric. In performance-based
earthquake engineering, explicit consideration of uncertainty is taken into account through
probabilistic evaluations of liquefaction susceptibility, triggering, and consequences combined
with a probabilistic evaluation of the ground motion hazard to produce estimates of liquefaction
risks that are rational and consistent. This type of approach was used to great effect in the design
of liquefaction mitigation measures for the TONEN tank farm complex in Tokyo, Japan, in the
1980s (see Box 9.1).
This chapter outlines the general approaches to performance-based evaluation and design
and presents recent application of the approaches. Appendix D provides details regarding the
implementation of performance-based design within the required probabilistic framework.
PERFORMANCE-BASED EVALUATION AND DESIGN 191

BOX 9.1
Performance-Based Design of the TONEN Tank Farm, Tokyo Bay, Japan

The TONEN tank farm complex is one of few published cases in which performance-based design
was applied to a project for which liquefaction was a major consideration (Baecher and Christian, 2003).
The refinery and tank farm sit on reclaimed land in a seismically active region in Tokyo Bay. The tanks
are separated into patios, each surrounded by firewalls intended to contain oil spilled from a single tank.
The principal risk of an off-site spill was foundation failure or deformation resulting as a consequence of
liquefaction that could cause firewall failure, large differential settlements of the tank bottoms, and tank
rupturing and leakage. Failure of a firewall around a ruptured tank or failure of multiple tanks within a
patio that leads to overtopping of the firewalls could result in oil spilling off-site into Tokyo Bay.
The site consists of approximately 8 meters of fill overlying recently deposited alluvial sands and
clays. The SPT blow counts are less than 10 blows/30 cm within the top ~5 meters, which indicates that
the site is highly liquefiable. The performance-based assessment of risk consisted of five steps:

1. Identification of factors influencing seismic performance of the storage tanks: frequency of


earthquakes, variability of soil properties, behavior of the tanks during earthquakes, performance
of the foundation soils during earthquakes, performance of the firewalls and revetments during
earthquakes, and performance of other site facilities designed to prevent spills.
2. Probabilistic seismic hazard analysis at the site to establish levels of ground shaking and their
respective exceedance probabilities.
3. Development of an event tree for a typical patio; the event tree was used to describe how various
facilities in a tank patio could interact to lead to system failure.
4. Analysis of the realizations identified in the event tree to develop a statistical sample of the
behavior of the site and facility.
5. Calculation of the probability of an off-site spill and the corresponding consequences of other
risks; the comparison and the annual expected monetary loss were used to judge the
acceptability of the liquefaction risk.
An important goal of the performance-based evaluation was to assess the cost- effectiveness of a
remediation strategy to lower the groundwater level beneath the facility. This remediation strategy would
not only reduce the thickness of saturated, liquefiable soil but also decrease the porewater pressure and
increase the effective vertical stress at depth and thereby decrease both the probability and the
consequences of liquefaction.
To understand the risk faced by the tank farm owners, the annual exceedance probability was
computed for different values of monetary losses. The annualized cost of failure due to earthquakes was
larger than was the annual cost of preventive measures, which led to the decision that the cost of
continuous lowering of the groundwater level was justified economically. An impervious slurry wall was
built around the site, pumps were installed, and a continuous program of pumping and monitoring of
groundwater levels was instituted. Although the 2011 Tohoku earthquake caused extensive liquefaction of
reclaimed land around Tokyo Bay (Yasuda et al., 2012), the TONEN tank farms experienced no damage.

APPROACHES TO PERFORMANCE-BASED EVALUATION

The performance-based approach can be used in many ways, most simply by implementing
them at the response level. Measures of system response (e.g., settlement, lateral spreading
displacement values) may be specified at various hazard levels or return periods. From the
response values, physical damage and loss are then inferred, often guided by judgement and
experience. A large uncertainty may be associated with loss estimates derived in this way. An
intermediate approach specifies performance related to damage limit states, which involves
comparison of the predicted and allowable responses (i.e., limit state). This approach involves
prediction of damage (e.g., minor cracking, number of pipe breaks per km) considering
192 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

uncertainties associated with the response and damage models. Loss is inferred from damage in
this approach, but there is less uncertainty in inferring loss from specific damage levels than
from specific response levels. The most complete approach defines performance in terms of
losses, including both loss of life and monetary loss; it involves the explicit modeling of
response, damage, and loss accounting for the uncertainties inherent in each. While requiring
more effort, loss level implementation provides the most accurate and least uncertain estimate of
loss (Kramer, 2014).

Response-Level Implementation

In practice, performance-based evaluations are usually implemented at the response level.


For liquefaction problems, this typically involves evaluation of the potential for triggering
followed by evaluation of the anticipated ground movements using a response model. To be
employed in the performance-based framework, both analyses must be performed using
probabilistic techniques. The probabilistic approaches described below illustrate the importance
of considering all levels of shaking that can lead to triggering and consequences of liquefaction
as well as considering the uncertainty in the predictions. These approaches allow development of
a response hazard curve that indicates how often, on average, different levels of response can be
expected. For example, a response hazard curve may provide the level of displacement that has a
2,475-year return period. Response hazard curves account for local seismicity (e.g., including all
expected earthquake magnitudes at all potential source-to-site distances); uncertainty in ground
motions given those magnitude and distance combinations; and uncertainties in prediction of
liquefaction triggering and movement of soils as a consequence of liquefaction movements (e.g.,
settlement, lateral spreading).
The level of response that has a 2,475-year return period will be different from the response
computed deterministically based on a 2,475-year ground motion. Basing design criteria on a
particular return period for liquefaction (Kramer and Mayfield, 2005; Kramer et al., 2006;
Franke et al., 2014) or on particular return periods of settlement or lateral spreading displacement
provides more uniform performance across different seismic environments than the current
practice of basing evaluations on deterministic factors of safety (FS) computed for a single
ground motion hazard level.
Fully probabilistic procedures for evaluating liquefaction triggering have been developed
(Marrone et al., 2003; Kramer and Mayfield, 2005, 2007; Juang et al. 2008; Baker and Faber,
2008; Mayfield et al. 2010; Franke et al., 2014). Kramer and Mayfield (2007) described a
probabilistic liquefaction hazard analysis procedure for evaluating liquefaction triggering that
generates FS hazard curves (see Figure 9.2) that account for all peak ground acceleration (PGA)
levels and all magnitudes that contribute to those PGA levels. These curves show the mean
annual rate of non-exceedance for different values of FS against triggering,  FS L . The reciprocal
of the value of  FS L for FSL = 1.0 is the return period of liquefaction (i.e., the average length of
time between occurrences of liquefaction of that element of soil). The hazard curves in Figure
9.2 were calculated for a hypothetical site profile placed in five cities with different seismic
hazards. The mean annual rates of non-exceedance are higher for the more seismically active
areas (e.g., California versus Tennessee), but not by a constant factor because the curves reflect
the differences in ground motion recurrence and the underlying magnitude contributions in
PERFORMANCE-BASED EVALUATION AND DESIGN 193

different areas. Kramer and Mayfield (2007) showed that application of conventional,
deterministic liquefaction hazard evaluation procedures used in practice (i.e., achieving a
minimum FS based on ground shaking from a single return period) produced very different
liquefaction hazards at the different sites, with the return period of liquefaction varying by a
factor of two or more. This result demonstrates that current deterministic procedures do not
provide similar levels of safety against liquefaction at sites in different seismic environments, as
quantified by the liquefaction hazard, and justifies movement toward the probabilistic approach.

FIGURE 9.2 Hazard curves for non-exceedance of a factor of safety (FS) against liquefaction at 6-m depth in
standard soil profile (inset) defined by Kramer and Mayfield (2007). The reciprocal of  FS L for FSL = 1.0 is the return
period of liquefaction for each location. SOURCE: Kramer, S.L., and R.T. Mayfield. 2007. The return period of
liquefaction. Journal of Geotechnical and Geoenvironmental Engineering 133(7):802–813. With permission from
ASCE. This material may be downloaded for personal use only. Any other use requires prior permission of the
American Society of Civil Engineers. This material may be found at
[http://ascelibrary.org/doi/abs/10.1061/%28ASCE%291090-0241%282007%29133%3A7%28802%29].

The calculation of a liquefaction FS hazard curve requires a cyclic resistance ratio curve
with an assigned probability of liquefaction. As noted in Chapter 4, the meaning of the
probabilities associated with various liquefaction curves in the published literature may be
ambiguous. In their original form, the curves may represent the probability of the observed data
given that liquefaction occurs rather than the probability of liquefaction given the observed data.
Therefore, users of probabilistic liquefaction curves should be sure that they understand the basis
for the curves they employ and that they are appropriate for the situation at hand (see Chapter 4,
Box 4.2).
Kramer and Huang (2010) used a response-level performance-based design framework with
a probabilistic settlement model (Huang, 2008) to compute hazard curves for post-liquefaction
(free-field) settlement. Figure 9.3 shows settlement hazard curves for a hypothetical site in
Seattle, Washington. Empirical lateral spreading models can be used to compute lateral
spreading displacement hazard curves in a slightly different manner. Franke and Kramer (2014)
developed a probabilistic version of the lateral spreading relationship developed by Youd and
colleagues (2002) and used it to compute displacement hazard curves for a generic site assumed
to be located in 10 U.S. cities with very different seismic environments (see Figure
9.4).
194 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Although each site was predicted to move approximately 30 cm under the loading associated
with the 475-year ground motion using the deterministic displacement prediction procedure
developed by Youd and colleagues (2002), the computed return periods from the fully
probabilistic analyses of these sites showed that the actual return periods for 30 cm of
displacement ranged from 77 to more than 2,300 years.
A conventional deterministic analysis, which ignores uncertainty in the displacement
prediction, implicitly assumes the return period of the response is the same as the return period
of the design ground motion (475 years in this case). When accounting for uncertainty in the
response through a fully probabilistic analysis, however, the return period for the response is not
equal to the return period of the ground motion. Given that there is uncertainty in the response
prediction, the probabilistic analysis provides the more accurate assessment of the response
return period.

FIGURE 9.3 Post-liquefaction settlement hazard FIGURE 9.4 Lateral spreading displacement hazard curves.
curves. SOURCE: Kramer, S.L., and Y.-M. Huang. SOURCE: Franke, K.W., and S.L. Kramer. 2014. Procedure
2010. Performance-based assessment of for the empirical evaluation of lateral spread displacement
liquefaction hazards. In Proceedings of the 9th U.S. hazard curves. Journal of Geotechnical and
National and 10th Canadian Conference on Geoenvironmental Engineering 140(1):110–120. With
Earthquake Engineering, Toronto, Canada. 10 pp. permission from ASCE. This material may be downloaded for
With permission from the Earthquake Engineering personal use only. Any other use requires prior permission of
Research Institute. the American Society of Civil Engineers. This material may be
found at
[http://ascelibrary.org/doi/abs/10.1061/%28ASCE%29GT.194
3-5606.0000969].

Damage-Level Implementation

A damage-level implementation of performance-based design considers the ground motion


hazard curve, a response model, and a damage model. The damage model is commonly
displayed through fragility curves, which predicts the probability of a damage state (e.g., minor
cracking) being exceeded given a response level (e.g., displacement). Attempts to explicitly
predict damage in the structural engineering discipline are more advanced than within
geotechnical engineering. Structural engineers often refer to maximum allowable damage as a
damage limit state.
PERFORMANCE-BASED EVALUATION AND DESIGN 195

Bird and colleagues (2005) performed structural analyses to evaluate damage to reinforced
concrete frame buildings from differential ground movements and developed fragility curves for
damage limit states while considering uncertainties in input parameters. To isolate the effects of
liquefaction, they used damage models to compare damage scenarios with and without ground-
failure-induced damage at a liquefiable site, and they noted the need for improved
characterization of the relationships between ground movement and structural damage.
Ledezma and Bray (2010) analyzed the potential damage to a five-span bridge supported by
piles that extended through liquefiable soils susceptible to lateral spreading. The response
models accounted for uncertainty in liquefaction triggering, uncertainty in the residual strength
of the liquefiable soil, and uncertainty in the lateral displacement prediction. They defined five
discrete damage states (none, small, moderate, large, and collapse) based on predicted
displacement. For a site located in Oakland, California, the computed probabilities of reaching
the five damage states are shown for two specific ground motion levels in Figure 9.6. Because
only two ground motion levels were considered, this approach is not considered fully
probabilistic, but the analysis shows that for a given ground motion level a range of damage
states is possible because of uncertainties in the response prediction and damage prediction.

(a) (b)
FIGURE 9.5 Fragility curves for damage to a specific class of reinforced concrete building for (a) horizontal
differential displacement, and (b) vertical differential displacement. Limit states 1-3 mark respective boundaries
between slight to moderate damage, moderate to extensive damage, and extensive to complete damage. SOURCE:
after Bird, J.F., Crowley, H., Pinho, R., and Bommer, J.J. (2005). “Assessment of building response to liquefaction-
induced differential ground deformation,” published in the Bulletin of the New Zealand Society for Earthquake
Engineering, Vol. 38, No.4, December 2005. Pp. 215-234.
196 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 9.6 Probabilities of reaching different damage states for bridge abutments at two hazard levels. Damage
states are none (N), small (S), moderate (M), large (L), and collapse (C). SOURCE: Ledezma, C., and J.D. Bray.
2010. Probabilistic performance-based procedure to evaluate pile foundations at sites with liquefaction-induced
lateral displacement. Journal of Geotechnical and Geoenvironmental Engineering 136(3):464–476. With permission
from ASCE. This material may be downloaded for personal use only. Any other use requires prior permission of the
American Society of Civil Engineers. This material may be found at
[http://ascelibrary.org/doi/abs/10.1061/%28ASCE%29GT.1943-5606.0000226].

Loss-Level Implementation

The most complete performance-based design procedure specifies performance in terms of


expected losses. Although such evaluations represent the most advanced application of
performance-based design, they are, for now, only practical for particularly large or particularly
important projects due to the significant work required to perform the analyses. Nevertheless,
they represent the future of earthquake engineering practice (Kramer, 2014).
Iai and colleagues (2008) described the design of a caisson quay wall in approximately 15
meters of water resting on liquefiable foundation soils and with liquefiable backfill. They
considered five options to improve the foundation soil and the backfill soil behind the quay wall
and developed fragility curves for damage state based on a series of parametric finite element
response analyses. Lifecycle costs for each of the five options were calculated based on a repair
cost of 1 million yen/m (approximately $3,000/ft) of wall length (see Figure 9.7). Construction
costs were similar for all the options, so the predicted direct losses differed only slightly.
However, Option B in Figure 9.7 turned out to be the clear choice for minimizing costs because
its indirect losses (i.e., those associated with loss of functionality) were lower.
Kramer and colleagues (2009) analyzed the same bridge considered by Ledezma and Bray
(2010). Using finite element analyses to predict response, the computed loss curves expressed as
the ratio of repair cost to replacement cost, termed the repair cost ratio, or RCR (see Figure 9.8a).
The losses were deaggregated to show the contributions of the various repair methods to the
predicted loss at a particular return period (see Figure 9.8b). Such deaggregations can provide an
engineer and owner with a useful indication of the origin of the losses.
PERFORMANCE-BASED EVALUATION AND DESIGN 197

Life-Cycle
Costs

Options
A: Foundation compaction only
Indirect Losses
B: Foundation cementation
C: Foundation and backfill compaction (1.8 m spacing)
D: Foundation & backfill compaction (1.6 m spacing)
Direct Losses
E: Foundation compaction & structural modification

Construction Costs

FIGURE 9.7 Illustration of life cycle costs for different soil improvement design options (after Iai, 2008). SOURCE:
Kramer, 2011.

(a) (b)
FIGURE 9.8 (a) Loss curve and (b) cost breakdown of 475-yr loss for five-span highway bridge. SOURCE: Modified
from Kramer, S.L., P. Arduino, and H.S. Shin. 2009. Development of performance criteria for foundations and earth
structures in Performance-Based Design in Earthquake Geotechnical Engineering: From Case History to Practice,
edited by T. Kokusho, Y. Tsukamoto, and M. Yoshimine, London, UK: Taylor & Francis. With permission from Taylor
& Francis Books UK.

Bradley and colleagues (2010) described a loss-level evaluation of a pile-supported bridge


underlain by liquefiable soil in New Zealand. Finite element analyses of the soil-foundation-
bridge system were used to predict system response related to nine response parameters,
including peak pile curvature, maximum abutment seating displacement, and approach
embankment settlement. Discrete damage states (e.g., cracking, yielding, and failure of piles)
were defined for each of the response parameters, and repair costs and repair durations were
defined, with uncertainties, for each of the discrete damage states. Figure 9.9 shows loss curves
for both the direct repair costs and the indirect costs associated with loss of functionality, and
these curves show that the costs associated with loss of functionality are far greater than are the
direct costs of the repairs themselves. Foundation compaction was accomplished by sand
compaction piles at indicated spacings.
198 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE 9.9 Loss curves from a performance-based evaluation of [a] pile-supported bridge. SOURCE: Bradley, B.A.,
M. Cubrinovski, R.P. Dhakal, G.A. MacRae. 2010. Probabilistic seismic performance and loss assessment of a
bridge-foundation-soil system. Soil Dynamics and Earthquake Engineering 30(5):395–411. With permission from
Elsevier.

FUTURE DEVELOPMENTS FOR PERFORMANCE-BASED EVALUATIONS

The calculations required for probabilistic performance evaluations may not be complicated,
but they are almost always voluminous. A full loss-level implementation requires integration
over all ground motion levels, response levels, and damage levels. For a given structure, detailed
loss analyses will recognize the existence of multiple foundation-structure components within a
structure; therefore, multiple response parameters and damage measures need to be computed. In
these cases, the between-component correlations must be determined and taken into account
while integrating over all components. Even the loss estimate may involve multiple parameters if
multiple components of loss (e.g., repair costs and loss of functionality) are considered. Because
fully probabilistic performance-based evaluations have been developed only recently, computer
programs and tools for their convenient and efficient use are in relatively early stages of
development. Programs that perform response-level implementations are currently available (see,
e.g., Huang, 2008; Franke et al., 2014), and programs for full loss-level implementations are in
development (see, e.g., Bradley, 2009; Silva et al., 2014). Additional development of these
programs and tools is needed for performance-based evaluation and design to be adopted widely
in engineering practice.
Fully probabilistic procedures show that uncertainties, whether in prediction of liquefaction
triggering or of consequences, significantly influence the calculated liquefaction hazards.
Improved response models are required that better quantify uncertainty and allow better
understanding of how much uncertainty can be reduced through such measures as increased
subsurface exploration, increased field and laboratory testing, more sophisticated analyses,
additional research, or other means. The ability to demonstrate the cost-effectiveness of such
measures is a major benefit of probabilistic performance-based liquefaction hazard evaluations.
Performance has several components, including physical damage, casualties, direct and
indirect costs, and loss of functionality, and it is viewed differently by different stakeholders.
Advances in performance-based procedures for liquefaction problems will require improved
understanding of the damage to different types of structures and facilities caused by different
PERFORMANCE-BASED EVALUATION AND DESIGN 199

levels of liquefaction-related ground deformation as well as improved understanding of the costs


and time requirements of repairing damage associated with liquefaction.
Data available today, in some cases, may not be adequate to quantify uncertainty and bias.
In such cases, subjective assessments may be necessary for some parameters. As performance-
based design becomes more widely used in geotechnical earthquake engineering, additional data
and experience will lead to more accurate quantification of uncertainties and less subjective
estimates of the probabilities of different levels of performance.
Performance-based approaches offer great opportunities to produce more reliable structures
and facilities and to use available resources more efficiently. Adoption of performance-based
engineering concepts into practice, however, will occur as engineers continue to refine their
thinking about uncertainty and risk. A fundamental shift from dependence on empirical
conventions and designs based on factors of safety toward design processes based on realistic
risk-based predictions of the response of engineered systems. This is a more scientifically
oriented approach that emphasizes accurate site characterization and performance prediction as
well as explicit consideration and quantification of uncertainties.
10
Recommendations

Major findings and conclusions of the committee that support the recommendations in this
chapter are summarized in the “Key Findings and Conclusions” boxes at the beginning of each
chapter. This chapter offers recommendations for improving engineering practice and scientific
understanding in regard to the triggering and consequences of liquefaction based on an
assessment of past and current research presented in the earlier chapters of this report. As such,
recommendations in this report reinforce old ideas, expand or move beyond current practice, or
offer new directions for research. Some of these recommendations can be implemented with
existing tools and are intended to help geotechnical engineers make better decisions in daily
practice. Others are directed at researchers and are intended to advance understanding of
liquefaction-related phenomena so that current approaches to assessing liquefaction triggering
and its consequences may be improved. Still other recommendations look beyond current
approaches toward new assessment methods. Developing and implementing these methods will
require collaboration among engineers, geologists, and others, and many would benefit from
community-based efforts. These will certainly require some level of funding. The organization
and funding of such efforts is beyond the scope of this report.
The recommendations are grouped by the main topics in the committee’s statement of task
(see Box 1.4):

(i) collecting, reporting, and assessing the sufficiency and quality of data;
(ii) addressing the spatial variability and uncertainty of these data; and
(iii) developing improved tools for assessing liquefaction triggering and its consequences.

Each boldfaced recommendation is followed by summaries of supporting evidence and, where


appropriate, examples of the improvements or information needed.

200
RECOMMENDATIONS 201

COLLECTING, REPORTING, AND ASSESSING DATA SUFFICIENCY AND


QUALITY

Recommendation 1. Establish curated, publicly accessible databases of relevant


liquefaction triggering and consequence case history data. Include case histories in which
soils interact with built structures. Document the case histories with relevant field,
laboratory, and physical model data. Develop the databases with strict protocols and
include indicators of data quality.

Field data from liquefaction case histories underpin the development, calibration, and
validation of predictive models for liquefaction triggering and its consequences. For maximum
benefit to the earthquake engineering community, case history databases need: (1) cases with
parameter values beyond the ranges found in current databases; (2) updating to include the
wealth of data available from recent earthquakes; (3) greater consistency with respect to
information included in the databases and greater transparency about differences in quality,
levels of detail, degree of vetting, and documentation; and (4) open access with capabilities for
searching.
Large amounts of liquefaction case history data collected since 1995 are not currently
included in the most widely used predictive models. Incorporating this information could
enhance the completeness of the available field data with regard to such parameters as tectonic
setting, earthquake magnitude, and soil type.
Current databases generally lack field case histories that represent behavior of liquefiable
soils

 at depths greater than about 15 meters beneath the ground surface and initial vertical
effective stresses larger than approximately 100 kPa (1 atm);
 subjected to relatively small and very large magnitude events (i.e., less than about M 5.9
and greater than approximately M 7.8);
 containing greater than 35% fines content;
 containing greater than 50% fines content and of low plasticity; and
 that are in sloping ground or are adjacent to free faces (i.e., may be subject to lateral
spreading) and have normalized standard penetration test (SPT) values (N1)60 (or (N1)60-cs
if adjusted clean-sand values are used) of greater than 15 blows per 30 cm or normalized
cone penetration test (CPT) resistance values qc1N (or qc1Ncs if adjusted clean-sand values
are used) greater than 85 atmospheres (8.6 MPa).

The databases are also unbalanced overall, being more heavily populated with cases where
liquefaction has been observed than where it has not occurred. Case histories where potentially
liquefiable soils have not liquefied can be as important as case histories where liquefaction has
occurred, and they are particularly important in establishing limits for liquefaction triggering and
its consequences with regard to magnitude, depth, initial static shear stress, grain size, and
plasticity. Future field studies should take particular notice of these data gaps and strive to fill
them with case histories that would influence the boundary between liquefaction and non-
liquefaction.
Current databases of consequences of liquefaction—especially of the residual shear strength
of liquefied soil, of lateral spreading displacements, and of the interaction between liquefied soils
202 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-INDUCED
SOIL LIQUEFACTION AND ITS CONSEQUENCES

and man-made structures—also need to be strengthened. This is particularly true for intermediate
density and dense soils where liquefaction effects may be subtle or possibly inconsequential:
namely, soils with a normalized CPT tip resistance, qc1-cs, greater than about 85 atm (8.6 MPa);
standardized and normalized SPT resistance, N1,60-cs, greater than approximately 15 blows per 30
cm; or a normalized shear wave velocity, VS1, greater than about 225 m/s. Collecting data from
sites that meet these criteria and that have been subjected to strong ground motions, even if there
are no observed effects of liquefaction, is a means to fill this data gap.
Few compilations of case histories document soil-structure interaction effects at sites that
have liquefied. Protocols should be developed for collecting data on soil-structure interaction
effects (e.g., the impact of lateral spreading on deep foundations and buried pipelines, bearing
failure of shallow foundations, the impact of preexisting structures on the amount of lateral
spreading), and more attention should be paid to collecting this type of data.
Relevant laboratory and physical model data can both augment field data in publicly
accessible, high-quality databases and help improve prediction models for conditions not
adequately constrained by case history data. Strict protocols for data included in the databases,
such as those being developed for the Next-Generation Liquefaction project (see Chapter 3, Box
3.3), need to be established and followed to ensure consistency of quality. Implementation of
Recommendation 1 cannot be fully realized without an adequately funded and coordinated effort.

Recommendation 2. Characterize locations with high probability of liquefaction and


establish them as field observatories for liquefaction triggering and its consequences.

Important gaps in case history databases could be filled by establishing liquefaction


observatories at sites that are well characterized, well instrumented, and strategically located in
areas where there is a high probability of earthquake-induced soil liquefaction in coming
decades. Detailed high-quality characterization of the pre-earthquake surface and subsurface
conditions at field observatories that then capture liquefaction events would strengthen the value
of recorded data by removing some uncertainty related to pre-liquefaction soil properties and
liquefaction-induced effects (e.g., changes in soil properties and ground displacements). Data
from these observatories can be used to calibrate and validate procedures for assessment of
liquefaction triggering and its consequences. Properly instrumented observatories may help fill
important data gaps, including behaviors associated with liquefaction at depth, liquefaction of
gravels and fine-grained soils, liquefaction on sloping ground and within and beneath
embankments, and soil-structure interaction effects in liquefied soil. Information from such
observatories could inform community-based collaborative efforts that will allow a more
complete understanding of liquefaction triggering and consequences.
Characterization data to be collected at observatory sites, where feasible, should include the
sedimentary and seismological history of the site, modifications made by humans, results of both
in situ and laboratory testing, lateral and vertical variability in liquefiable soils, and surface and
subsurface evidence of previous liquefaction. Site characterization also needs to include the
collection of high-quality geographically referenced and orthorectified aerial and satellite images
as well as high-precision topographic data (e.g., from Light Detection and Ranging surveys).
These data need to be updated periodically to account for geomorphic or anthropogenic changes
that may occur prior to recorded earthquake activity. Instrumentation should include strong
motion instruments at the surface and at depth (ideally above and below liquefiable layers) and
pore-pressure transducers within liquefiable soil strata and at other strategic locations in the soil
RECOMMENDATIONS 203

profile; inclinometers in areas subject to potential lateral spreading; settlement probes; and
appropriate instrumentation (e.g., strong motion recorders, strain gages, inclinometers) of
structural elements of on-site facilities (e.g., piles, foundation slabs, building superstructures, and
bridge decks).
Several free-field borehole arrays in the United States and elsewhere (see Box 3.3) meet the
minimum requirements for observatories for evaluation of both empirical and numerical
liquefaction triggering models. More sites with this type of instrumentation are needed to
account for the diverse geologic, seismologic, and geotechnical conditions where liquefaction is
possible. Given the low probability of a triggering event at any specific site, having multiple
instrumented sites will increase the likelihood of capturing a triggering event.
Few case histories that include detailed measurements of soil-foundation-structure
interaction effects on liquefiable ground are available. More sites at which instrumentation is
installed to capture such soil-structure interaction effects in liquefied soils are needed. A cost-
effective technique to obtain useful data may be to place new liquefaction field observatories at
already well-characterized sites (e.g., sites instrumented as part of a strong motion
instrumentation program). There are challenges associated with implementing and funding such
a program, but the California Strong Motion Instrumentation Program,1 established primarily to
provide data on the seismic response of structures, is one model of how such observatories can
be funded and maintained.

Recommendation 3. Use data from the cone penetration test (CPT) for field-based
estimates of liquefaction resistance where feasible. If the standard penetration test (SPT) is
used for this purpose, make hammer energy measurements. Supplement field-based
estimates with other methods as appropriate to characterize the site.

CPT soundings offer advantages over other methods of estimating liquefaction resistance in
detecting thin layers that may affect liquefaction triggering and subsequent pore-pressure
redistribution. CPT results are less dependent on the equipment operator or setup than most other
in situ test methods, and the CPT can be performed with relative speed and economy. The CPT
can also provide a cost-effective means to measure shear wave velocity (VS) that can enhance
site characterization and liquefaction assessment. CPT equipment is available that can collect
samples, but this method is not generally available, and the technology at the time of this writing
is not mature. Using seismic piezocone penetration testing (CPTu) should be considered because
the additional information from the CPTu can provide insight to soil layering, soil dilatancy, and
groundwater conditions, and the additional information from the seismic cone can be used for
other purposes (e.g., soil layering, site response analyses, and as an alternative method to
evaluate liquefaction triggering). There are sites, however, where CPT soundings are not
feasible, such as those with gravelly or very dense soils. In such cases, other techniques for
assessing the liquefaction resistance of the soil, including the SPT, alternative means to collect
Vs measurements, the Becker Penetration Test (BPT), and large diameter dynamic cone
penetration testing, have proven useful. If the BPT is used, an instrumented BPT rig in which
energy delivered to the sampler is measured (i.e., the BPTi), should be used.
The SPT is still widely used in practice and has a well-defined role in cases where direct
measurement of fines content is needed or is desirable to reduce uncertainties (e.g., for
1
See www.consrv.ca.gov/cgs/smip.
204 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-INDUCED
SOIL LIQUEFACTION AND ITS CONSEQUENCES

interpretation of CPT and Vs data). SPT blow counts may also be useful as a supplemental
technique for assessing soil resistance in dense soils and at depths where the CPT cannot
penetrate. Nonetheless, there are several operational factors for which corrections are difficult to
make. The most variable of these, and that which often requires the largest correction factor, is
hammer energy. If the SPT blow count is to be used to assess resistance of a soil to liquefaction,
SPT hammer energy measurements are needed to reduce the uncertainty associated with SPT
blow count values. The SPT setup should conform to conditions for which no correction factor is
needed (e.g., standard borehole diameter, sampler configuration). Rotary wash borings for SPT
testing are more reliable than are hollow stem auger borings, as SPT blow counts recorded in
hollow stem auger borings below the water table are particularly susceptible to error and should
be carefully evaluated for indications of borehole disturbance (i.e., abnormally low blow count
values).
Use of multiple methods for any assessment may help to constrain uncertainties. The source
of any discrepancies in assessments should be identified to inform professional judgement of the
design engineer and to allow the engineer to decide which methods are most reliable in the given
circumstances.

Recommendation 4. When refining or developing new empirical relationships for use in


liquefaction analyses, incorporate unbiased estimates for input parameters; identify and
quantify when possible the uncertainty associated with those estimates; and use soil
mechanics principles, seismologic principles, and experimental data to extrapolate beyond
ranges in which field data constrain the empirical relationships.

Relationships for analysis of liquefaction triggering and consequences based on case history
data include cyclic resistance curves developed for simplified stress-based analysis of triggering,
relationships between post-liquefaction shear strength and penetration resistance, and equations
for evaluating the levels of liquefaction-related lateral spreading. Empirical procedures generally
yield consistent estimates of liquefaction hazards for conditions within the ranges that are
constrained by the field data. Databases of field case histories, however, do not include data for
the full range of conditions over which a liquefaction assessment may be required (e.g., depths of
about 15 m or greater, gravelly or silty soils, earthquakes larger than magnitude 8, presence of
static shear stress from a topographic slope). Empirical procedures that employ a polynomial
functional form based solely on the best statistical fit to available data may yield unrealistic
results when used to make predictions beyond the limits of the data. In contrast (as discussed in
Chapter 4), extrapolation using functional forms based on experimental data trends (e.g., from
simple shear and triaxial tests and from centrifuge testing) and on principles of soil mechanics,
wave propagation theory, or seismology, as appropriate, provide a more appropriate means of
extrapolating beyond the limits of the field data. Developers of empirical procedures need to
state clearly the range over which their procedures are constrained by the field data and the
method used to extrapolate beyond that range.
Those developing or refining empirical relationships should attempt to use approaches that
minimize bias, rather than simplified equations with a built-in bias, when quantifying the
adjustment factors, correlations, and parameter relationships used in the relationship. Developers
also need to identify and, if possible, to quantify the associated uncertainty with each adjustment
factor, empirical correlation, and parameter relationship. This will avoid the compounding of
uncertainties and will facilitate assessment of the overall uncertainty associated with the method.
RECOMMENDATIONS 205

Employing adjustment factors, empirical correlations, and parameter relationships with built-in
bias when developing a liquefaction triggering or consequence assessment method compounds
uncertainties and makes it difficult, if not impossible, to assess the overall uncertainty associated
with the method—an important step in liquefaction assessments. Assessing the overall
uncertainty is an important consideration when recommending a factor of safety (FS) to use in a
liquefaction triggering or consequence assessment.

ADDRESSING THE SPATIAL VARIABILITY AND UNCERTAINTY OF DATA

Recommendation 5. Use geology to improve the geotechnical understanding of case


histories and project sites, particularly where potentially liquefiable soils vary in thickness,
continuity, and engineering properties.

Current site assessment practice tends to focus on the engineering characteristics of


subsurface materials. Their geologic context, however, is basic to assessing liquefaction hazards
both in case histories and in project work. Opportunities to use geology more effectively include
incorporating descriptions of sedimentary geology in protocols for site assessment of case
histories and in site characterization for project work. In case histories, inferences about which
deposits liquefied, and which did not, could include characterization of the vertical and lateral
variability of postulated critical layers (deposits inferred to have liquefied) and of subsurface
paths of fluid escape that dikes and sills can mark. Such geologic investigations and
characterizations could aid interpretation of case histories where the sedimentary geometry is
complex, as is commonly the case on alluvial fans and in river meander belts, and they could be
extended to project work where detailed understanding of liquefaction potential is warranted by
threats to life and property. Understanding the lateral continuity and variability of liquefiable
layers can be particularly important in assessing the potential for lateral spreading.
Geologic understanding of surficial deposits also underpins regional maps of liquefaction
hazard that can serve as screening tools in project work. These regional maps—sometimes
prepared by state agencies—are becoming increasingly quantitative through studies that relate
surficial geologic units to probabilistic estimates of liquefaction potential. Further improvements
to the maps could include representing subsurface stratigraphy where liquefaction may occur in
stratigraphic units that widely underlie those mapped at the surface.
A database of geologic evidence for liquefaction in the central and eastern United States
could be expanded worldwide, and it could provide information on the sedimentary environment,
depositional age, and tectonic setting to guide evaluation of the influence of these factors on
liquefaction potential. This effort, in turn, might enable case history databases to be made more
complete by including data on geologic evidence for liquefaction in earthquakes in the 19th
century and before. Extending temporal coverage of liquefaction case histories in this manner
could be particularly helpful in areas where major earthquakes recur at intervals of hundreds or
thousands of years and where there are few case histories of any age.
These suggested efforts are meant to strengthen use of geology in assessments of
liquefaction triggering and consequences. They would supplement current best practices that
include accurate description, sound interpretation, and qualified professional judgment regarding
tectonics, landforms, regional geology, sedimentology, groundwater hydrology, and past
performance of analogous sites.
206 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-INDUCED
SOIL LIQUEFACTION AND ITS CONSEQUENCES

Recommendation 6. Implement simplified stress-based methods for liquefaction triggering


in a manner consistent with how they were developed. Avoid using techniques and
adjustment factors from one variant of a method with other variants. Consider using more
than one simplified method when making a liquefaction triggering assessment.

The simplified stress-based approach to liquefaction triggering analysis has served the
profession for more than 40 years and, despite shortcomings, will continue to be a mainstay of
earthquake engineering practice for the near future. Several variants of the simplified stress-
based method have been developed. Each variant has associated adjustment factors, empirical
correlations, and parameter relationships that are used during its implementation. If the
adjustment factors, empirical correlations, and parameter relationships from one variant of the
simplified stress-based procedure are applied to another variant (e.g., the mixing and matching of
factors, correlations, and relationships) errors are likely to be introduced, and both bias and
uncertainty will be added to the liquefaction triggering assessment.
Each of the variants of the simplified stress-based method, for example, includes a means of
assessing the earthquake-induced shear stress with depth as the stress reduction coefficient, rd.
The location of the triggering curve in a particular variant of the simplified method depends upon
the method used to assess rd. Therefore, using rd from one variant of the simplified method (e.g.,
the Idriss and Boulanger, 2008, variant) in conjunction with the triggering curve from another
variant (e.g., the Cetin et al., 2004, variant) will introduce unquantified (and unquantifiable)
errors into the triggering assessment.
Similarly, it is tempting to assume that, rather than using a simplified equation for rd to
assess the earthquake-induced shear stress, using a more precise analytical method for assessing
earthquake-induced shear stress (e.g., a site response analysis) will increase the accuracy of the
analysis. In instances where the simplified equation for rd incorporates biased estimates of
earthquake-induced shear stress (see, e.g., Idriss and Boulanger, 2008), however, using a
different method to compute the shear stresses will introduce additional and unquantifiable
uncertainty into the liquefaction assessment.
There are significant epistemic (modeling) uncertainties associated with all variants of the
simplified stress-based approach to liquefaction triggering assessment. Situations may warrant
the use of more than one variant, such as where analysis results indicate a marginal FS against
triggering, and where the consequences of triggering are considered to be unacceptable (e.g.,
require costly remediation or threaten life safety). A more sophisticated liquefaction assessment
may be warranted if the results of the triggering assessment in these situations are contradictory
or marginal. Multiple variants of the simplified method are also warranted when the method is
applied beyond the bounds where method relationships are constrained by field data (see
Recommendation 1 for these limits).
As stated previously, using multiple methods may help constrain uncertainties among test
methods. The sources of discrepancies among test method results should be identified. It is then
up to the professional judgment of the investigator to determine how results should be weighed,
or whether the use of more advanced methods may be warranted.
RECOMMENDATIONS 207

Recommendation 7. In developing methods to evaluate liquefaction triggering and its


consequences, explicitly incorporate uncertainties from field investigations, laboratory
testing, numerical modeling, and the impact of the local site conditions on the earthquake
ground motions.

It is especially important for both the developer of methods and the practicing engineer to
consider uncertainty in procedures used for assessment of liquefaction triggering and its
consequences. All aspects of liquefaction assessment—from characterizing the site and site-
specific ground motions to assessing the severity of liquefaction consequences—are fraught with
uncertainty. The uncertainties in field data and analytical and experimental results need to be
stated in the form of error bounds, standard deviations, bias, or other statistically appropriate
measures.
Most field investigations, laboratory test programs, and analysis procedures do not explicitly
address uncertainty. Descriptions of procedures to evaluate liquefaction do not always state
clearly where and what the uncertainties are and typically do not address how to incorporate
them into the liquefaction assessment. Furthermore, while many factors used for liquefaction
assessment require correction, the correction factors (e.g., corrections applied to SPT blow
counts, magnitude scaling factor) are themselves uncertain. The uncertainty in the correction
factor is rarely accounted for, and in some cases, it may introduce even more uncertainty in the
results of the analysis than that associated with the original measurement (see Box 4.3).
Uncertainty in a liquefaction consequence or triggering assessment increases rapidly as the
assessment method moves beyond the range within which it is constrained by field data. Because
the data ranges over which many assessment procedures are applicable may not be adequately
identified, users of the procedures may be tempted to extrapolate beyond the ranges for which
the procedures are valid without considering the increased uncertainty. Good engineering
practice dictates that any extrapolation beyond existing data be supported with lab testing or
other sources of information. Liquefaction assessment procedures need to clearly identify the
range over which they are constrained by both field and laboratory test data. Users need either to
refrain from extrapolating beyond this range or to explicitly identify when they have done so and
qualify the results accordingly.

Recommendation 8. Refine, develop, and implement performance-based approaches to


evaluating liquefaction, including triggering, the geotechnical consequence of triggering,
structural damage, and economic loss models to facilitate performance-based evaluation
and design.

Earthquake engineering is, in general, moving toward performance-based design, and


liquefaction issues will need to be addressed within that framework. Regional variations in
seismicity, coupled with the significant uncertainties in earthquake loading and liquefaction
resistance, warrant probabilistic characterization for the evaluation of both liquefaction
triggering potential and the consequences of liquefaction. Quantifying the uncertainties
associated with existing and new procedures is necessary to facilitate risk-based approaches to
liquefaction triggering and consequence assessment. Geotechnical engineers need to acquire and
document data to support quantification of uncertainties and to develop efficient computational
tools that make risk-based calculations practical.
208 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-INDUCED
SOIL LIQUEFACTION AND ITS CONSEQUENCES

Probabilistic liquefaction triggering and consequence evaluation procedures can be


convolved with the results of a probabilistic seismic hazard analysis (PSHA) to produce fully
probabilistic liquefaction hazard analyses (PLHA). A PLHA considers all levels of shaking
(rather than just that associated with a single return period), all contributing magnitudes (rather
than just the mean or modal magnitude) at all return periods, and the uncertainty in ground
motion at all levels to provide a complete and consistent indication of liquefaction hazards across
different seismic environments. To accomplish this, computational tools need to be integrated
with PSHA tools, making certain they allow weighted contributions from multiple predictive
models to account for epistemic uncertainty.
The geotechnical community needs to develop PLHA to meet the growing demand from the
greater earthquake engineering community for risk-based liquefaction assessment as part of the
trend toward performance-based design. The results of a PLHA should be expressed in the form
of a response curve that indicates how often different levels of response (e.g., liquefaction
triggering, lateral spreading displacement, settlement) can be expected to occur. In a
performance-based framework, the results of this type of PLHA can be convolved with damage
and loss models to estimate the risk associated with liquefaction. Risk can be expressed as a risk
curve that indicates how often different levels of loss can be expected to occur. Losses include
loss of life, loss of functionality, and direct and indirect economic consequences.
Implementation of performance-based design often requires the use of advanced
computational methods. Keeping in mind the need to use more than one strong ground motion
time history, the amount of computation required for even a single time history analysis, and the
need to conduct sensitivity analyses regarding modeling assumptions, a balance must be struck
between the complexity of the numerical model, the detail to which the site is or can be
characterized, and computational efficiency. The balance to be struck depends on the specifics of
the problem: for instance, what is of primary concern. No general guidance can be provided on
how to strike this balance; therefore, the site investigation and numerical analysis plans and
analyses results should be peer reviewed by an engineer experienced in this type of analysis, as
recommended by the ASCE (2010) for site-specific seismic hazard assessments.

IMPROVING TOOLS FOR ASSESSING


LIQUEFACTION TRIGGERING AND ITS CONSEQUENCES

Recommendation 9. Use experimental data and fundamental principles of seismology,


geology, geotechnical engineering, and engineering mechanics to develop new analytical
techniques, screening tools, and models to assess liquefaction triggering and post-
liquefaction consequences.

The simplified methods for liquefaction triggering and consequence assessment that are
typically employed in practice today are limited in their predictive capabilities. This limitation
also exists with the more sophisticated empirical and analytical methods used for liquefaction
assessment. Furthermore, most available methods for assessment of liquefaction triggering and
consequence assessment are not compatible with probabilistic characterization of the seismic
loading that has become standard in engineering practice. New approaches to liquefaction
assessment are needed that would go beyond improving empirical methods (Recommendation 4).
Ideally, these new approaches will:
RECOMMENDATIONS 209

(a) relate quantitatively liquefaction hazards to geologic units at appropriate depths;


(b) increase predictive capabilities in both evaluation of triggering and consequences;
(c) be compatible with probabilistic characterization of seismic loading;
(d) be consistent with performance-based assessments of liquefaction consequences; and
(e) be readily accessible to practicing engineers.

To relate liquefaction hazards quantitatively to geologic units at appropriate depths,


geotechnical engineers will have to work with engineering geologists to characterize the depth-
dependent distribution and properties of geologic units. To increase the accuracy of predictions,
new approaches need to be based on sound science (e.g., geology, seismology, and physics) and
fundamental principles of engineering mechanics and geotechnical engineering. The approaches
need to be consistent with patterns of behavior identified through analysis of field case histories
and laboratory and physical model test data. Strain-based and energy-based approaches are two
avenues of development that merit consideration.
A next logical step in model development is to introduce approaches compatible with
probabilistic seismic hazard characterization that can include or be integrated with liquefaction
consequence assessment. Rigorous validation of these models using case history data and
development of implementation approaches that make the models accessible to practitioners is
needed to facilitate their adoption in engineering practice. Development and adoption in practice
of such integrated and validated models would be consistent with trends in other areas of modern
earthquake engineering practice.

Recommendation 10. Develop and validate computational models for liquefaction analyses.
Use laboratory and physical model tests at different spatial scales and case histories to
provide insight into fundamental soil behavior and to validate the application of
constitutive models to boundary-value problems.

While constitutive and computational models are intimately linked, computational models
do not always accurately reproduce behavior across laboratory-test, physical-test, and field
scales. Some limitations of computational modeling are related to the inability of the models to
capture complex material behavior under cyclic loading, including the intimately coupled
deformation and flow patterns before and after triggering. Such patterns may induce changes to
solid-to-viscous-liquid behavior transformations, localized and large deformations, and porosity
redistribution in the ground, all of which can only be predicted faithfully with improved
computational models. Models based on discrete element mechanics offer a promising avenue
for modeling mechanisms at the grain level, but they remain computationally burdensome.
Further developments are needed to enhance the accuracy and computational efficiency of
discrete methods and to couple the discrete methods with simulations across scales. Mesh-free
methods, such as the material point method, offer great potential for modeling large strain flow
behavior. Rigorous validation procedures that distinguish between accuracy of the computational
method and the complexity of the problems for which the predictive model has been validated
are required for all computational procedures, regardless of the computational approach and
method of analysis they employ.
210 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-INDUCED
SOIL LIQUEFACTION AND ITS CONSEQUENCES

Recommendation 11. Conduct fundamental research on the stress, strain, and strength
behaviors of soils prior to and after liquefaction triggering; devise new laboratory and
physical model experimental techniques to aid development of constitutive models of those
behaviors.

Fundamental research on liquefaction-related phenomena, both theoretical and


experimental, is still required to advance the state of the art and practice. The stress-strain
behavior of soils prior to and following liquefaction triggering is complex and still not fully
understood. No available constitutive model captures all of the relevant soil behavior within a
rigorous physics-based framework. In addition to changes in stiffness and strength of the soil
caused by earthquake-induced excess porewater pressure, there are also changes in soil fabric
and porosity. Needed research in these areas includes research on dilation that occurs in soils
prior to and after liquefaction triggering, the solid–to-viscous-liquid behavior transformation that
can accompany triggering, and the mechanisms that control the post-triggering fabric
degradation that influences stress-strain behavior of soils. In particular, the mechanisms of post-
triggering particle rearrangement and reconsolidation and the effects of relative density and
consolidation stress on these mechanisms are not well understood and require further research.
Understanding the mechanisms of post-liquefaction soil behavior is fundamental to
understanding the dependency of post-liquefaction residual strength on consolidation stress; this
includes the potential for a threshold value of soil resistance (e.g., normalized SPT blow count or
CPT tip resistance) above which residual shear strength increases so precipitously that flow
sliding and lateral spreading are not of concern.

IMPROVING RESEARCH AND PRACTICE

Today’s empirical approaches to liquefaction assessment reflect a half century of


development and refinement. Society has benefited from widespread acceptance and application
of these approaches, but now those approaches to liquefaction assessment are changing in
response to new data, new analytical methods, and recent developments in earthquake
engineering. The preceding recommendations stress the importance of implementing existing
methods in accordance with their stated procedures, using high-quality data that depict spatial
variations at a site, and making that high-quality data available publicly after careful review and
vetting. The recommendations also emphasize the development of fully PLHAs that can be used
in a performance-based design framework. In the meantime, predictive models and tools
consistent with fundamental principles of dynamics and soil mechanics need to be developed to
meet the need of extrapolating beyond the ranges of existing data.
Advances such as these will require a concerted effort not only by researchers and
practitioners but also by facility owners and stakeholders at large. Researchers need to continue
to collect high-quality field case history data supplemented with laboratory and model test data
as well as to develop and validate new, more accurate models to assess earthquake-induced soil
liquefaction. Practitioners need to make greater efforts to understand the limitations of their
choices of characterization and assessment approaches, and they need to better understand the
specific geologic and tectonic controls on liquefaction hazards at their sites. Facility owners need
to contribute to this effort by supporting the establishment of public case history databases and
well-characterized and instrumented field observatories for study of liquefaction effects.
RECOMMENDATIONS 211

Perhaps most importantly, liquefaction assessment needs to expand beyond empirically


based methods with new methods that are informed by more data, by quantification of
uncertainties, and by fundamental scientific and engineering principles. To be adopted by the
engineering community, any new method needs to be responsive to practitioner needs. In
particular, the

(a) fundamental basis and applicability of the model needs to be understandable;


(b) outputs from the model need to be translatable to the problems of interest;
(c) limitations of the model need to be clearly defined;
(d) model needs to be readily implementable by practicing engineers;
(e) methods must be economical (i.e., not require excessive effort); and the
(f) methods must keep pace with the evolution of earthquake engineering practice toward
performance-based design.

Moving along these recommended paths forward will fundamentally improve prediction of
earthquake-induced soil liquefaction and its consequences. Reliable predictions will provide
greater protection of life safety, of built infrastructure, and of the economy by reducing adverse
economic, environmental, and social impacts of liquefaction.
References

AASHTO (American Association of State Highway and Transportation Officials). 2014. AASHTO Guide
Specifications for LRFD Seismic Bridge Design with 2012 and 2014 Interim Revisions.
Washington, DC: American Association of State Highway and Transportation Officials.
Available at http://app.knovel.com/hotlink/toc/id:kpAASHTO53/aashto-guide-specifications;
accessed 15 November 2016.
Abdoun, T. 1997. Modeling of seismically induced lateral spreading of multilayer sand deposit and its
effects on pile foundations. Ph.D. diss. Troy, NY: Department of Civil Engineering, Rensselaer
Polytechnic Institute. 630 pp.
Abrahamson, N.A., B.A. Bolt, R.B. Darragh, J. Penzien, and T.B. Tsai. 1987. The SMART 1
accelerograph array (1980–1987): A review. Earthquake Spectra 3:263–287.
ALA (American Lifelines Alliance). 2001. Seismic Fragility Formulations for Water Systems. Available
at http://americanlifelinesalliance.com/pdf/Part_1_Guideline.pdf; accessed 18 February 2016.
Alarcon-Guzman, A., G.A. Leonards, and J.L. Chameau. 1988. Undrained monotonic and cyclic strength
of sands. Journal of Geotechnical Engineering 114(10):1089–1108.
Ambraseys, N.N. 1988. Engineering seismology. Earthquake Engineering and Structural Dynamics
17:1–105.
Amos, C.B., A.T. Lutz, A.S. Jayko, S.A. Mahan, G.B. Fisher, and J.R. Unruh. 2013. Refining the
southern extent of the 1872 Owens Valley earthquake rupture through paleoseismic investigations
in the Haiwee area, Southeastern California. Bulletin of the Seismological Society of America
103:1022–1037, doi:10.1785/0120120024.
Anderson, D.G., G.R. Martin, I. Lam, and J.N. Wang. 2008. Seismic Analysis and Design of Retaining
Walls, Buried Structures, Slopes and Embankments, Recommended Specifications, Commentaries
and Example Problems. National Cooperative Highway Research Program Report
611.Washington, DC: Transportation Research Board. 148 pp.
Anderson, D.G., S. Seungcheol, and S.L. Kramer. 2011. Observations from nonlinear, effective-stress
ground motion response analyses following the AASHTO guide specifications for LRFD seismic
bridge design. Transportation Research Record: Journal of the Transportation Research Board
2251:114–154, DOI: 10.3141/2251-15.
Andrade, J.E. 2009. A predictive framework for liquefaction instability. Géotechnique 59(8):673–682.

212
REFERENCES 213

Andrade, J.E., and R.I. Borja. 2007. Modeling deformation banding in dense and loose fluid-saturated
sands. Finite Elements in Analysis and Design 43:361–383.
Andrade, J.E., and X. Tu. 2009. Multiscale framework for behavior prediction in granular media.
Mechanics of Materials 41(6):652–669.
Andrade, J.E., A.M. Ramos, and A. Lizcano. 2013. Criterion for flow liquefaction instability. Acta
Geotechnica 8:525–535.
Andrews, D.C.A., and G.R. Martin. 2000. Criteria for liquefaction of silty soils. Paper 0312 in
Proceedings of the 12th World Conference on Earthquake Engineering, 30 January–4 February
2000, Auckland, New Zealand. Upper Hutt, N.Z.: New Zealand Society for Earthquake
Engineering.
Andrus, R.D. 1986. Subsurface investigations of a liquefaction induced lateral spread Thousand Springs
Valley, Idaho: Liquefaction recurrence and a case history in gravel. Master’s thesis. Provo, UT:
Department of Civil Engineering, Brigham Young University. 117 pp. Available at
https://ceen.et.byu.edu/sites/default/files/snrprojects/245-ronald_d_andrus-1986-tly.pdf; accessed
18 February 2016.
Andrus, R.D., and K.H. Stokoe II. 1997. Liquefaction resistance based on shear wave velocity. Pp. 89–
128 in Proceedings of the NCEER Workshop on Evaluation of Liquefaction Resistance of Soils,
5–6 January 1996, Salt Lake City, Utah. Buffalo, NY: National Center for Earthquake
Engineering Research.
Andrus, R.D., and K.H. Stokoe II. 2000. Liquefaction resistance of soils from shear-wave velocity.
Journal of Geotechnical and Geoenvironmental Engineering 126(11):1015–1025.
Andrus, R.D., and T.L. Youd. 1987. Subsurface investigation of a liquefaction-induced lateral spread,
Thousand Springs Valley, ID. Geotechnical Laboratory Miscellaneous Paper GL-87-8.
Washington, DC: U.S. Army Corps of Engineers.
Andrus, R.D., and T. L. Youd. 1989. Penetration tests in liquefiable gravels. Pp. 679–682 in Proceedings
of the 12th International Conference on Soil Mechanics and Foudation Engineering, Rio de
Janeiro, Brazil. Boca Raton, FL: CRC Press.
Andrus, R.D., K.H. Stokoe, J.A. Bay, and T.L. Youd. 1992. In situ Vs of gravelly soils which liquefied.
Pp. 1447–1452 in Proceedings of the Tenth World Conference on Earthquake Engineering, 19–
24 July 1992, Madrid, Spain. Rotterdam, the Netherlands: A.A. Balkema.
Andrus, R.D., K.H. Stokoe, R.M. Chung, and C.H. Juang. 2003. Guidelines for Evaluating Liquefaction
Resistance Using Shear Wave Velocity Measurements and Simplified Procedures. NIST GCR 03-
854. Gaithersburg, MD: National Institute of Standards and Technology.
Andrus, R.D., H. Hayati, and N. Mohanan. 2009. Correcting liquefaction resistance of aged sands using
measured to estimated velocity ratio. Journal of Geotechnical and Geoenvironmental
Engineering 135(6):735–744.
Arango, I., M.R. Lewis, and C. Kramer. 2000. Updated liquefaction potential analysis eliminates
foundation retrofitting of two critical structures. Soil Dynamics Earthquake Engineering 20:17–
25.
Archuleta, R.J., S.H. Seale, P.V. Sangas, L.M. Baker, and S.T. Swain. 1992. Garner Valley downhole
array of accelerometers: Instrumentation and preliminary data analysis. Bulletin of Seismological
Society of America 82:1592–1621.
Arduino, P. 2014. Consequences of Liquefaction Analytical Models. Presentation to the Committee on
State of the Art and Practice in Earthquake Induced Soil Liquefaction Assessment, 11 March,
Tempe, Arizona State University. National Academies of Sciences, Engineering, and Medicine.
Arias, A. 1970. A measure of earthquake intensity. Pp. 438–483 in Seismic Design for Nuclear Power
Plants, edited by R.J. Hansen. Cambridge, MA: MIT Press.
Armstrong, R.J., and E.J. Malvick. 2014. Comparison of liquefaction susceptibility criteria. Pp. 29–37 in
Dams and Extreme Events—Reducing Risk of Aging Infrastructure under Extreme Loading
Conditions: 34th Annual USSD Conference, San Francisco, California, April 7-11, 2014. Denver,
CO: United States Society on Dams.
214 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Armstrong, R.J., and E.J. Malvick. 2015. Practical considerations in the use of liquefaction susceptibility
criteria. Earthquake Spectra 32(3), doi:10.1193/071114EQS100.
Arulanandan, K. 2008. Soil Structure: In Situ Properties and Behavior. Colombo, Sri Lanka:
Samayawardhana Printers.
Arulanandan, K., and R. Scott. 1993. Project VELACS—Control test results. Journal of Geotechnical
Engineering 119(8):1276–1292.
Arulmoli, K., and K. Arulanandan. 1994. In Situ Electrical Method for Evaluation of Stress Ratio
Required to Cause Liquefaction and Dynamic Modulus. Symposium on Dynamic Geotechnical
Testing II, 27–28 January, Reno, Nevada. American Society for Testing and Materials.
Arulmoli, K., K. Arulanandan, and H.B. Seed. 1985. New method for evaluating liquefaction potential.
Journal of the Geotechnical Engineering Division 111(1):95–114.
ASCE (American Society of Civil Engineers). 1998. Seismic Guidelines for Ports. Edited by S.D.
Werner. Technical Council on Lifeline Earthquake Engineering (TCLEE) Monograph No. 12.
Reston, VA: American Society of Civil Engineers. 377 pp.
ASCE. 2010. Minimum design loads for buildings and other structures. Standards ASCE/SEI 7-10.
Reston, VA: American Society of Civil Engineers, Structural Engineering Institute. 636 pp.
ASCE. 2014. Seismic Design of Piers and Wharves. Standards ASCE/COPRI 61-14. Reston, VA:
American Society of Civil Engineers.
Ashford, S.A., and K.M. Rollins. 2002. TILT: Treasure Island liquefaction test final report. Rep. No.
SSRP-2001/17. Department of Structural Engineering, University of California, San Diego.
Ashford, S.A., T. Juirnarongrit, T. Sugano, and M. Hamada. 2006. Soil–pile response to blast-induced
lateral spreading. I: Field Test. Journal of Geotechnical and Geoenvironmental Engineering
132(2):152–162.
Ashford, S., R.W. Boulanger, and S. Brandenberg. 2011. Recommended Design Practice for Pile
Foundations in Laterally Spreading Ground. PEER 2011/04. Pacific Earthquake Engineering
Research Center, College of Engineering, University of California, Berkeley.
ASTM (American Society for Testing and Materials). 2011. Standard Practices for Cycle Counting in
Fatigue Analysis. ASTM Standard E1049-85(2011)e1. West Conshohocken, PA: ASTM
International, doi:10.1520/E1049-85R11E01.
ASTM Committee D-18 on Soil and Rock. 2010. Standard test methods for liquid limit, plastic limit, and
plasticity index of soils. West Conshohocken, PA: ASTM International.
ASTM Committee D-18 on Soil and Rock. 2011. Standard practice for classification of soils for
engineering purposes (unified soil classification system). West Conshohocken, PA: ASTM
International.
ASTM International. 2011a. Standard Test Method for Standard Penetration Test (SPT) and Split-Barrel
Sampling of Soils. ASTM D1586-11. West Conshohocken, PA: ASTM International.
ASTM International. 2011b. Standard Practice for Determining the Normalized Penetration Resistance
of Sands for Evaluation of Liquefaction Potential. ASTM-D6066-11. West Conshohocken, PA:
ASTM International.
ASTM International. 2012. Standard Test Method for Electronic Friction Cone and Piezocone
Penetration Testing of Soils. ASTM D5778-12. West Conshohocken, PA: ASTM International.
ASTM International. 2014a. Standard Test Methods for Downhole Seismic Testing. ASTM D7400-14.
West Conshohocken, PA: ASTM International.
ASTM International. 2014b. Standard Test Methods for Crosshole Seismic Testing. ASTM D4428. West
Conshohocken, PA: ASTM International.
ASTM International. 2015. Standard Practice for Sampling of Soil Using the Hydraulically Operated
Stationary Piston Sampler. ASTM D6519-15. West Conshohocken, PA: ASTM International.
Atkinson, G.M., J.J. Bommer, and N.A. Abrahamson 2014. Alternative approaches to modeling epistemic
uncertainty in ground motions in probabilistic seismic‐hazard analysis. Seismological Research
Letters 85(6), doi:10.1785/0220140120.
REFERENCES 215

Bachman, R., D. Nyman, K. Bhushan, E.V. Leyendecker, and L. Lister. 2007. Draft seismic design,
guidelines and data submittal requirements for LNG facilities. Prepared for the Federal Energy
Regulatory Commission. Available at http://www.ferc.gov/industries/gas/indus-act/lng/lng-seis-
guide.pdf; accessed 18 February 2016.
Baecher, G.B., and J.T. Christian. 2003. Reliability and Statistics in Geotechnical Engineering.
Chichester, U.K.: John Wiley & Sons, Inc.
Baise, L.G., R.B. Higgins, and C.M. Brankman. 2006. Liquefaction hazard mapping - Statistical and
spatial characterization of susceptible units. Journal of Geotechnical and Geoenvironmental
Engineering 132:705–715, doi:10.1061/(ASCE)1090-0241(2006)132:6(705).
Baker, J. 2011. Conditional mean spectrum: Tool for ground-motion selection. Journal of Structural
Engineering 137:322–331, doi:10.1061/(ASCE)ST.1943-541X.0000215.
Baker, J., and M. Faber. 2008. Liquefaction Risk Assessment Using Geostatistics to account for Soil
Spatial Variability. Journal on Geotechnical and Geoenvironmental Engineering 134(1):14–23.
Bardet, J.P., and F. Liu. 2009. Motions of gently sloping ground during earthquakes. Journal of
Geophysical Research 114(1):125–165.
Bardet, J.P., T. Tobita, N. Mace, and J. Hu. 2002. Regional modeling of liquefaction-induced ground
deformation. Earthquake Spectra 18(1):19–46.
Bartlett, S.F., and T.L. Youd. 1992. Empirical analysis of horizontal ground displacement generated by
liquefaction-induced lateral spreads. Technical Report NCEER-92-0021. Buffalo, NY: National
Center for Earthquake Engineering Research.
Bartlett, S.F., and T.L. Youd. 1995. Empirical prediction of liquefaction-induced lateral spread. Journal
of Geotechnical and Geoenvironmental Engineering 121(4):316–329.
Bathe, K.J. 1996. Finite Element Procedures in Engineering Analysis. Upper Saddle River, NJ: Prentice
Hall Inc.
Baxter, C.P.D., A.S. Bradshaw, R.A. Green, and J.-H. Wang. 2008. Shear wave velocity correlation for
liquefaction resistance of silt. Journal of Geotechnical and Geoenvironmental Engineering
134(1):37–46.
Bazant, Z.P., and L. Cedolin. 2010. Stability of Structures: Elastic, Inelastic, Fracture and Damage
Theories. Singapore: World Scientific Publishing Co.
Beaty, M., and P. Byrne. 1998. An effective stress model for predicting liquefaction behavior of sand. Pp.
766–777 in Geotechnical Earthquake Engineering and Soil Dynamics III, Vol. 1¸ edited by P.
Dakoulas, M. Yegian, and R. D. Holtz. ASCE Geotechnical Special Publication No. 75. Reston,
VA: American Society of Civil Engineers.
Beaty, M., and P. Byrne. 2011. UBCSAND Constitutive Model Version 904aR: Documentation Report.
Available at http://www.itasca-
udm.com/media/download/UBCSand/UBCSAND_UDM_Documentation.pdf; accessed 29
November 2016.
Beaty, M., and V.G. Perlea. 2011. Several Observations on Advanced Analyses with Liquefiable
Materials. 21st Century Dam Design-Advances and Adaptations, 31st Annual USSD Conference.
11–15 April, San Diego, California. United States Society on Dams.
Belytschko, T., W.K. Liu, and B. Moran. 2000. Nonlinear Finite Elements for Continua and Structures.
Chichester, U.K.: John Wiley & Sons, Inc.
Bennett, M.J., and J.C. Tinsley III. 1995. Geotechnical Data from Surface and Subsurface Samples
Outside of and Within Liquefaction-Related Ground Failures Caused by the October 17, 1989,
Loma Prieta Earthquake, Santa Cruz and Monterey Counties, California. U.S. Geological Survey
Open-File Report 95-663. Available at http://pubs.usgs.gov/of/1995/0663/report.pdf; accessed 18
February 2016.
Berger, M.J., and P. Colella. 1989. Local adaptive mesh refinement for shock hydrodynamics. Journal of
Computational Physics 82:64–84.
Berrill, J.B., and R.O. Davis. 1985. Energy dissipation and seismic liquefaction of sands: Revised model.
Soils and Foundations 25(2):106–118
216 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Biondi, G., E. Cascone, and M. Maugeri. 2004. Number of uniform stress cycles equivalent to seismic
loading. Pp. 705–712 in Proceedings of the 11th International Conference on Seismic Design and
Earthquake Engineering and 3rd International Conference on Earthquake Geotechnical
Engineering, Vol. 2, University of California, Berkeley, edited by D. Doolin et al. Singapore:
Stallion Press.
Biot, M.A. 1941. General theory of three dimensional consolidation. Journal of Applied Physics
12(2):155–164.
Biot, M.A. 1956a. Theory of propagation of elastic waves in a fluid-saturated porous solid. I. Low-
frequency range. The Journal of the Acoustical Society of America 28(2):168–178.
Biot, M. A. 1956b. Theory of propagation of elastic waves in a fluid-saturated porous solid. II. Higher
frequency range. The Journal of the Acoustical Society of America 28(2):179–191.
Bird, J.F., H. Crowley, R. Pinho, and J.J. Bommer. 2005. Assessment of building response to
liquefaction-induced differential ground deformation. Bulletin of the New Zealand Society for
Earthquake Engineering 38(4):215–234.
Bird, J.F., J.J. Bommer, H. Crowley, and R. Pinho. 2006. Modelling liquefaction-induced building
damage in earthquake loss estimation. Soil Dynamics and Earthquake Engineering 26(1):15–30.
Bishop, A.W. 1959. The principle of effective stress. Teknisk Ukeblad 106(39):859–863.
Bommer, J.J., and F. Scherbaum. 2008. The use and misuse of logic trees in probabilistic seismic hazard
analysis. Earthquake Spectra 24:997–1009.
Bonilla, L.F. 2001. NOAH: User’s manual. Institute for Crustal Studies, University of California, Santa
Barbara. 38 pp. Available online at
http://indico.ictp.it/event/7615/session/3/contribution/17/material/0/0.pdf; accessed 29 November
2016.
Bonilla, L.F., R.J. Archuleta, D. Lavallee. 2005. Hysteretic and dilatant behavior of cohesionless soils and
their effects on nonlinear site response: Field data observations and modeling. Bulletin of the
Seismological Society of America 95(6):2373–2395, doi: 10.1785/0120040128.
Bonilla, L., K. Tsuda, N. Pulido, J. Régnier, and A. Laurendeau. 2011. Nonlinear site response evidence
of K-NET and KiK-net records from the 2011 off the Pacific coast of Tohoku Earthquake. Earth
Planets Space 63(7):785–789.
Borja, R.I. 2000. A finite element model for strain localization analysis of strongly discontinuous fields
based on standard Galerkin approximation. Computational Methods in Applied Mechanics and
Engineering 190(11):1529–1549.
Borja, R.I. 2006. Condition for liquefaction instability in fluid saturated granular soils. Acta Geotechnica
1(4):211–224.
Boulanger, R.W. 1990. Liquefaction behavior of saturated cohesionless soils subjected to uni-directional
and bi-directional static and cyclic loads. Ph.D. diss., University of California, Berkeley.
Boulanger, R.W. 2003a. Relating K to a relative state parameter index. Journal of Geotechnical and
Geoenvironmental Engineering 129(10):770–773.
Boulanger, R.W. 2003b. High overburden stress effects in liquefaction analyses. Journal on Geotechnical
and Geoenvironmental Engineering 129(12):1071–1082.
Boulanger, R.W. 2010. A Sand Plasticity Model for Earthquake Engineering Applications. Report No.
UCD/CGM-10-01. Center for Geotechnical Modeling, Department of Civil and Environmental
Engineering, University of California, Davis. 77 pp.
Boulanger, R.W., and I.M. Idriss. 2004a. State normalization of penetration resistance and the effect of
overburden stress on liquefaction resistance. Pp. 484–491 in Proceedings of the 11th
International Conference on Seismic Design and Earthquake Engineering and 3rd International
Conference on Earthquake Geotechnical Engineering, Vol. 1, University of California, Berkeley,
edited by D. Doolin et al. Singapore: Stallion Press.
Boulanger, R.W., and I.M. Idriss. 2004b. Evaluating the potential for liquefaction or cyclic failure of silts
and clays. Report No. UCD/CGM-04/01. Center for Geotechnical Modeling, Department of Civil
and Environmental Engineering, University of California, Davis.
REFERENCES 217

Boulanger, R.W., and I.M. Idriss. 2006. Liquefaction susceptibility criteria for silts and clays. Journal of
Geotechnical and Geoenvironmental Engineering 132(11):1413–1426.
Boulanger, R.W., and I.M. Idriss. 2008. Closure to Liquefaction Susceptibility Criteria for Silts and
Clays. Journal of Geotechnical and Geoenvironmental Engineering. 134(7):1027-1028.
Boulanger, R.W., and I.M. Idriss. 2014. CPT and SPT based liquefaction triggering procedures. Report
No. UCD/CGM-14/01. Center for Geotechnical Modeling, Department of Civil and
Environmental Engineering, University of California, Davis. 134 pp.
Boulanger, R.W., and I.M. Idriss. 2015. Magnitude scaling factors in liquefaction triggering procedures.
Soil Dynamics and Earthquake Engineering 79:296–303.
Boulanger, R.W., and R.B. Seed. 1995. Liquefaction of sand under bi-directional monotonic and cyclic
loading. Journal of Geotechnical Engineering 121(12):870–878.
Boulanger, R.W., and K. Ziotopoulou. 2012. PM4SAND (Version 2):A sand plasticity model for
earthquake engineering applications. Report No. UCD/CGM-12/01. Center for Geotechnical
Modeling, University of California, Davis. 96 pp.
Boulanger, R.W., and K. Ziotopoulou. 2013. Formulation of a sand plasticity plane-strain model for
earthquake engineering applications. Soil Dynamics and Earthquake Engineering 53:254–267.
Boulanger, R.W., and K. Ziotopoulou. 2015. PM4Sand (Version 3):A sand plasticity model for
earthquake engineering applications. Report No. UCD/CGM-15/01. Center for Geotechnical
Modeling, Department of Civil and Environmental Engineering, University of California, Davis.
112 pp.
Boulanger, R.W., R.B. Seed, C.K. Chan, H.B. Seed, and J. Sousa. 1991. Liquefaction behavior of
saturated sands under unidirectional and bi-directional monotonic and cyclic simple shear
loading. Geotechnical Engineering Rep. No. UCB/GT/91-08. University of California, Berkeley.
Boulanger, R.W., M.W. Meyers, L.H. Mejia, and I.M. Idriss. 1998. Behavior of a fine-grained soil during
the Loma Prieta earthquake. Canadian Geotechnical Journal 35:146–158.
Boulanger, R.W., B.L. Kutter, S.J. Brandenberg, P. Singh, and D. Chang. 2003. Pile Foundations in
Liquefied and Laterally Spreading Ground during Earthquake: Centrifuge Experiments &
Analyses. Report No. UCD/CGM-03/01. Center for Geotechnical Modeling, Department of Civil
and Environmental Engineering, University of California, Davis.
Boulanger, R.W., D. Chang, U. Gulerce, S. Brandenberg, and B.L. Kutter. 2005. Evaluating pile pinning
effects on abutments over liquefied ground. Pp. 306–318 in Seismic Performance and Simulation
of Pile Foundations in Liquefied and Laterally Spreading Ground. ASCE Special Geotechnical
Publication No. 145. Reston, VA: American Society of Civil Engineers,
doi:10.1061/40822(184)25.
Boulanger, R.W., D.W. Wilson, and I.M. Idriss. 2012. Examination and reevaluation of SPT-based
liquefaction triggering case histories. Journal of Geotechnical and Geoenvironmental
Engineering 138(8):898–909.
Bradley, B.A. 2009. User manual for SLAT: Seismic loss assessment tool version 1.14. Research Report
2009-01. Christchurch, N.Z.: Department of Civil Engineering, University of Canterbury. 94 pp.
Bradley, B.A. 2013. Site-specific and spatially distributed estimation of ground motion intensity in the
2010-2011 Canterbury earthquakes. Soil Dynamics and Earthquake Engineering 61-62:83–91.
Bradley, B.A., M. Cubrinovski, R.P. Dhakal, G.A. MacRae. 2010. Probabilistic seismic performance and
loss assessment of a bridge-foundation-soil system. Soil Dynamics and Earthquake Engineering
30(5):395–411.
Brandenberg, S.J., R.W. Boulanger, B.L. Kutter, and D. Chang. 2005. Behavior of pile foundations in
laterally spreading ground during centrifuge tests. Journal of Geotechnical and
Geoenvironmental Engineering 131(11):1378–1391.
Brankman, C.M. and L.G. Baise. 2008. Liquefaction susceptibility mapping in Boston, Massachusetts.
Engineering and Environmental Geoscience 14(1):1–16.
Bray, J.D., and S. Dashti. 2014. Liquefaction-induced building movements. Bulletin of Earthquake
Engineering 12(3):1129–1156.
218 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Bray, J.D., and D. Frost. 2010. Geo-Engineering Reconnaissance of the February 27, 2010 Maule, Chile
Earthquake. GEER Association Report No. GEER-022. Geotechnical Estrem Events
Reconnaissance Association.
Bray, J.D., and R.B. Sancio. 2006. Assessment of the liquefaction susceptibility of fine grained soils.
Journal of Geotechnical and Geoenvironmental Engineering 132(9):1165–1177.
Bray, J.D., and T. Travasarou. 2007. Simplified procedure for estimating earthquake-induced deviatoric
slope displacements. Journal of Geotechnical and Geoenvironmental Engineering 133(4):381–
392.
Bray, J.D, R.B. Sancio, T. Durgunoglu, A. Onalp, T.L. Youd, J.P. Stewart, R.B. Seed, O.K. Cetin, E. Bol,
M.B. Batuary, C. Christensen, and T. Karadayilar. 2004a. Subsurface characterization at ground
failure sites in Adapazari, Turkey. Journal of Geotechnical and Geoenvironmental Engineering
130(7):673–685.
Bray, J.D., R.B. Sancio, M.F. Riemer, and T. Durgunoglu. 2004b. Liquefaction susceptibility of fine-
grained soils. Pp. 655–662 in Proceedings of 11th International Conference on Soil Dynamics
and Earthquake Engineering and 3rd International Conference on Earthquake Geotechnical
Engineering, edited by D. Doolin et al. Singapore: Stallion Press.
Bray, J.D., M. Cubrinovski, J. Zupan, and M. Taylor. 2014. Liquefaction effects on buildings in the
central business district of Christchurch. Earthquake Spectra 30(1):85–109.
Budnitz, R.J., G. Apostolakis, D.M. Boore, L.S. Cluff, K.J. Coppersmith, C.A. Cornell, and P.A. Morris.
1997. Recommendations for Probabilistic Seismic Hazard Analysis: Guidance on Uncertainty
and the Use of Experts. NUREG/CR-6372. Washington, DC: U.S. Nuclear Regulatory
Commission.
Buscarnera, G., and A. Whittle. 2013. Model prediction of static liquefaction: influence of initial state on
potential instabilities. Journal of Geotechnical and Geoenvironmental Engineering 139(3):420–
432.
Byrne, P. 1991. A cyclic shear-volume coupling and pore-pressure model for sand. Pp. 46–55 in
Proceedings: Second International Conference on Recent Advances in Geotechnical Earthquake
Engineering and Soil Dynamics, 11–15 March, St. Louis, Missouri. Paper No. 1.24.
Byrne, P.M., S.S. Park, M. Beaty, M.K. Sharp, L. Gonzalez, and T. Abdoun. 2004. Numerical modeling
of liquefaction and comparison with centrifuge tests. Canadian Geotechnical Journal 41(2):193–
211.
California Geological Survey. 2004. Recommended criteria for delineating seismic hazard zones in
California. Special Publication 118. Sacramento: California Department of Conservation. 12 pp.
California Geological Survey. 2008. Guidelines for Evaluating and Mitigating Seismic Hazards in
California. Special Publication 117A. Sacramento: California Department of Conservation. 98
pp. Available at http://www.conservation.ca.gov/cgs/shzp/webdocs/Documents/sp117.pdf;
accessed 18 February 2016.
Calvetti, F., and R. Nova. 2004. Micromechanical approach to slope stability analysis. In Degradations
and instabilities in geomaterials, edited by F. Darve and I. Vardoulakis. CISM International
Centre for Mechanical Sciences 461:235–254.
Cao, Z., T. Youd, and X. Yuan. 2013. Chinese dynamic penetration test for liquefaction evaluation in
gravelly soils. Journal of Geotechnical and Geoenvironmental Engineering 139(8):1320–1333.
Carter, D.P., and H.B. Seed. 1988. Liquefaction potential of sand deposits under low levels of excitation.
UCB/EERC-88/11. Earthquake Engineering Research Center, University of California, Berkeley.
315 pp.
Carter, W.L., R.A. Green, B.A. Bradley, and M. Cubrinovski. 2014. The influence of near-fault motions
on liquefaction triggering during the Canterbury earthquake sequence. Pp. 57–68 in Soil
Liquefaction during Recent Large-Scale Earthquakes, edited by R.P. Orense, I. Towhata, and N.
Chouw. Leiden, the Netherlands: CRC Press.
Castro, G. 1969. Liquefaction of sands. Harvard Soil Mechanics Series 87. Cambridge, MA: Harvard
University.
REFERENCES 219

Castro, G. 1987. On the behaviour of soils during earthquake liquefaction. Pp. 169–204 in Soil Dynamics
and Liquefaction. Amsterdam, the Netherlands: Elsevier.
Castro, G. 1991. Determination of in-situ undrained steady state strength of sandy soils and seismic
stability of tailings dams. Pp. 111–113 in Proceedings of the 9th Panamerican Conference on
Soil Mechanics, Soc. Chilena de Mecanica de Suelos, Santiago, Chile.
Castro, G., and S.J. Poulos. 1977. Factors affecting liquefaction and cyclic mobility. Journal of the
Geotechnical Engineering Division 106(GT6):501–506.
Castro, G., and J. Troncoso. 1989. Effects of 1985 Chilean earthquake on three tailing dams. Pp. 35–59 in
as Jornadas Chilenas de Sismología e Ingeniería Antisísmica, 7-11 de Agosto de 1989, Santiago,
Chile.
Castro, G., S. Poulos, and F. Leathers. 1985. Re‐examination of slide of lower San Fernando dam.
Journal of Geotechnical Engineering 111:1093–1107, doi:10.1061/(ASCE)0733-
9410(1985)111:9(1093).
Castro, G., R.B. Seed, T.O. Keller, and H.B. Seed. 1992. Steady-state strength analysis of lower San
Fernando Dam slide. Journal of Geotechnical Engineering 118(3):406–427.
Ceccato, F., A. Rohe, and P. Simonini. 2014. Simulation of slope failure experiment with the material
point method. In Proceedings of Incontro Annuale dei Ricercatori di Geotecnica, Chieti, Italy.
Cetin, K.O., and H.T. Bilge. 2012. Cyclic large strain and induced pore pressure models for saturated
clean sands. Journal of Geotechnical and Geoenvironmental Engineering 138(3):309–323.
Cetin, K.O., and H.T. Bilge. 2014. Stress scaling factors for seismic soil liquefaction engineering
problems: A performance-based approach. In Perspectives on Earthquake Geotechnical
Engineering, edited by A. Ansal and M. Sakr. Basel, Switzerland: Springer.
Cetin, K.O., and H.T. Bilge. 2015. Stress Scaling Factors for Seismic Soil Liquefaction Engineering
Problems: A Performance-Based Approach. Pp 113–139 in Perspectives on Earthquake
Geotechnical Engineering, edited by A. Ansal and Muhamed Sakr. Geotechnical, Geological and
Earthquake Engineering 37. Basel, Switzerland: Springer International. doi:10.1007/978-3-319-
10786-8_5.
Cetin, K.O., and R.B. Seed. 2004. Nonlinear shear mass participation factor (rd) for cyclic shear stress
ratio evaluation. Soil Dynamics and Earthquake Engineering 24(2):103–113.
Cetin, K.O., R.B. Seed, R.E.S. Moss, A.K. Der Kiureghian, K. Tokimatsu, L.F. Harder, and R.E. Kayen.
2000. Field Performance Case Histories for SPT-Based Evaluation of Soil Liquefaction
Triggering Hazard. Geotechnical Engineering Research Report No. UCB/GT-2000/09.
Geotechnical Engineering, Department of Civil Engineering, University of California, Berkeley.
Cetin, K.O., R.B. Seed, A.K. Der Kiureghian, K. Tokimatsu, L.F. Harder, Jr., R.E. Kayen, and R.E.S.
Moss. 2004. Standard penetration test-based probabilistic and deterministic assessment of seismic
soil liquefaction potential. Journal of Geotechnical and Geoenvironmental Engineering
130(12):1314–1340.
Cetin, K.O., H.T. Bilge, J. Wu, A.M. Kammerer, and R.B. Seed. 2009a. Probabilistic models for the
assessment of cyclically induced reconsolidation (volumetric) settlements. Journal of
Geotechnical and Geoenvironmental Engineering 135(3):387–398.
Cetin, K.O., H.T. Bilge, J. Wu, A.M. Kammerer, and R.B. Seed. 2009b. Probabilistic models for cyclic
straining of saturated clean sands. Journal of Geotechnical and Geoenvironmental Engineering
135(3):371–386.
Chang, B.J., and T.C. Hutchinson. 2013. Experimental evaluation of p-y curves considering development
of liquefaction. Journal of Geotechnical and Geoenvironmental Engineering 139 (4):577 –586.
Chang, S.E. 2000. Disasters and transport systems: Loss, recovery and competition at the Port of Kobe
after the 1995 earthquake. Journal of Transport Geography 8(1):53-65.
Chang, W.-J., E.M. Rathje, K.H. Stokoe II, and B. R. Cox. 2004. Evaluation of effectiveness of
prefabricated drains in liquefiable sand. Soil Dynamics and Earthquake Engineering 24(9–
10):623–731.
220 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Chang, W.-J., Rathje, E.M., Stokoe, K. H., and Hazirbaba, K. 2007. In situ pore pressure generation
behavior of liquefiable sand. Journal of Geotechnical and Geoenvironmental Engineering
133:921–931.
Charlie, W.A., P.J. Jacobs, and D.O. Doehring. 1992. Blast-induced liquefaction of an alluvial sand
deposit. Geotechnical Testing Journal 15(1):14–23.
Chleborad, A.F., and R.L. Schuster. 1998. Ground failure associated with the Puget Sound region
earthquakes of April 13, 1949 and April 29, 1965. Pp. 373–439 in Assessing earthquake hazards
and reducing risk in the Pacific Northwest, Vol. 2, edited by A.M. Rogers, T.J. Walsh, W.J.
Kockelman, and G.R. Priest. U.S. Geological Survey Professional Paper 1560. Available at
http://pubs.usgs.gov/pp/p1560/; accessed 18 February 2016.
Christoffersen, J., M.M. Mehrabadi, and S. Nemat-Nasser. 1981. A micromechanical description of
granular material behavior. Journal of Applied Mechanics 48(2):339–344.
Chou, J.C., Kutter B.L., Travasarou T., and Chacko J.M. 2011. Centrifuge modeling of seismically
induced uplift for the BART Transbay Tube. Journal of Geotechnical and Geoenvironmental
Engineering 137(8):754-765.
Chung, J., and J. Rogers. 2011. Simplified method for spatial evaluation of liquefaction potential in the
St. Louis area. Journal of Geotechnical and Geoenvironmental Engineering 137(5):505–515.
City of Los Angeles. 2014. Letter to Geology and Soils Firms Practicing in the City of Los Angeles.
Department of Building and Safety, Los Angeles, California, 16 July.
CNS (National Standard of the People’s Republic of China). 2001. Code for seismic design of buildings.
National Standard GB 50011-2001. Beijing: China Building Industry Press.
Coduto, D.P., M.-C. Yeung, and W.A. Kitch. 2011. Geotechnical engineering: Principles and practices.
Upper Saddle River, NJ: Pearson.
Cook, B.K., D.R. Noble, and J.R. Williams. 2004. A direct simulation method for particle-fluid systems.
Engineering Computations 21:151–168.
Cornell, C.A., and H. Krawinkler. 2000. Progress and challenges in seismic performance assessment.
PEER News, 1–3 April.
Cortright, C.J. 1975. Effects of the San Fernando earthquake on the Van Norman Reservoir complex. Pp.
395–406 in San Fernando, California Earthquake of 9 February 1971, edited by Gordon B.
Oakeshott. Bulletin 196. Sacremento: California Department of Conservation, Division of Mines
and Geology.
Cox, B.R., K.H. Stokoe, and E.M. Rathje. 2009. An in-situ test method for evaluating the coupled pore
pressure generation and nonlinear shear modulus behavior of liquefiable soils. Geotechnical
Testing Journal 32(1):11–21.
Cox, S.C., H.K. Rutter, A. Sims, M. Manga, J. J. Wier, T. Ezzy, P.A. White, T.W. Horton, and D. Scott.
2012. Hydrological effects of the Mw 7.1 Darfield (Canterbury) earthquake, 4 September 2010,
New Zealand. New Zealand Journal of Geology and Geophysics 55(3):231-247.
Cramer, C.H., G.J. Rix, K. Tucker. 2008. Probabilistic liquefaction hazard maps for Memphis, Tennessee.
Seismological Research Letters 79(3):416–423.
Cubrinovski, M., and K. Ishihara. 1998. State concept and modified elastoplasticity for sand modeling.
Soils Found 38 (4):213–225.
Cubrinovski, M., K. Ishihara, and F. Tanizawa. 1996. Numerical simulation of the Kobe Port Island
liquefaction. Paper No. 330 in 11 WCEE: Eleventh World Conference on Earthquake
Engineering, 23–28 June, Acapulco, Mexico. Oxford, UK: Elsevier Science Ltd. 8 pp.
Cubrinovski, M., J.D. Bray, M. Taylor, S. Giorgini, B. Bradley, L. Wotherspoon, and J. Zupan. 2011. Soil
liquefaction effects in the central business district during the February 2011 Christchurch
earthquake. Seismological Research Letters 82(6):893–904.
Cubrinovski, M., K. Robinson, M. Taylor, M. Hughes, and R. Orense. 2012. Lateral spreading and its
impacts in urban areas in the 2010-2011 Christchurch earthquakes. New Zealand Journal of
Geology and Geophysics 55:255–269, doi:10.1080/00288306.2012.699895.
REFERENCES 221

Cubrinovski, M., M.W. Hughes, B.A. Bradley, J. Noonan, R. Hopkins, S. McNeill, and G. English. 2014.
Performance of horizontal infrastructure in Christchurch City through the 2010-2011 Canterbury
earthquake sequence. Civil & Natural Resources Engineering Research Report 2014-02. N.Z:
University of Canterbury. 129 pp.
Cundall, P.A., and O.D.L. Strack. 1979. A discrete numerical model for granular assemblies.
Géotechnique 29(1):47–65.
Dafalias, Y.F., and M.T. Manzari. 2004. Simple plasticity sand model accounting for fabric change
effects. Journal of Engineering Mechanics 130(6):622–634.
Daniel, C.R., J.A. Howie, and A. Sy. 2003. A method for correlating large penetration test (LPT) to
standard penetration test (SPT) blow counts. Canadian Geotechnical Journal 40:66–77.
Danisch, L.A., M.S. Lowery-Simpson, and T. Abdoun. 2004. Shape-acceleration measurement device and
method. Patent application. Available at http://www.google.com/patents/US7296363; accessed 18
February 2016.
Dashti, S., and J. D. Bray. 2013. Numerical simulation of building response on liquefiable sand. Journal
of Geotechnical and Geoenvironmental Engineering 139(8):1235–1249.
Davies, M.P., and R.G. Campanella. 1994. Selecting design values of undrained strength for cohesionless
soils. Pp. 176–186 in Proceedings of the 47th Canadian Geotechnical Conference, Vol. I,
Halifax, Nova Scotia: Bitech Publishers.
Davis, R.O., and J.B. Berrill. 1982. Energy dissipation and seismic liquefaction in sands. Earthquake
Engineering and Structural Dynamics 10:59–68.
Davis, R.O., and J.B. Berrill. 2001. Pore pressure and dissipated energy in earthquakes – Field
verification. Journal of Geotechnical and Geoenvironmental Engineering 127(3):269–274.
De Alba, P., C.K. Chan, and H.B. Seed. 1975. Determination of soil liquefaction characteristics by large-
scale laboratory tests. Report Number UCB/EERC-75/14. Earthquake Engineering Research
Center, University of California, Berkeley. 173 pp.
De Alba, P., H.B. Seed, and C.K. Chan. 1976. Sand liquefaction in large-scale simple shear tests. Journal
of the Soil Mechanics and Foundations Division 102(GT9):909–927.
De Alba, P.A., H.B. Seed, E. Retamal, and R. B. Seed. 1988. Analyses of dam failures in 1985 Chilean
earthquake. Journal of Geotechnical Engineering 114(12):1414–1434.
Deierlein, G.G., H. Krawinkler, and C.A. Cornell. 2003. A framework for performance-based earthquake
engineering. In Proceedings of 2003 Pacific Conference on Earthquake Engineering. 8 pp.
DeJong, J., M. Ghafghazi, A. Sturm, R. Armstrong, A. Perez, and C. Davis. 2014. A new Instrumented
Becker Penetration Test (iBPT) for improved characterization of gravelly deposits within and
underlying dams. The Journal of Dam Safety 12(2):9–20.
Desai, C.S., and H.J. Siriwardane. 1984. Constitutive Laws for Engineering Materials: With Emphasis on
Geologic Materials. Englewood Cliffs, NJ: Prentice Hall.
Dickenson, S. 2005. Recommended Guidelines for Liquefaction Evaluations using Ground Motions from
Probabilistic Seismic Hazard Analyses. Report to the Oregon Department of Transportation.
Available at http://www.oregon.gov/odot/hwy/bridge/docs/bddm/pdfs/psha.pdf; accessed 18
February 2016.
Dixit, J., D.M. Dewaikar, and J.S. Jangrid. 2012. Assessment of liquefaction potential index for Mumbai
City. Natural Hazards and Earth Science Systems 12:2759–2768.
Dobry, R. 1985. Unpublished data, personal files of R. Dobry, Rensselaer Polytechnic Institute, Troy,
NY. Figure included in NRC 1985.
Dobry, R., and T. Abdoun. 2011. An investigation into why liquefaction charts work: A necessary step
toward integrating the states of art and practice. Pp. 13–44 in Proceedings of the 5th International
Conference on Earthquake Geotechnical Engineering, 10–13 January, Santiago, Chile. Ishihara
Lecture.
Dobry, R., and T. Abdoun. 2015. Cyclic shear strain needed for liquefaction triggering and assessment of
overburden pressure factor Kσ. Journal of Geotechnical and Geoenvironmental Engineering
141(11), doi:10.1061/(ASCE)GT.1943-5606.0001342, 04015047.
222 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Dobry, R., and R.S. Ladd. 1980. Discussion to “Soil liquefaction and cyclic mobility evaluation for level
ground during earthquakes,” by H.B. Seed and “Liquefaction potential: Science versus practice,”
by R.B. Peck. Journal of the Geotechnical Engineering Division 106(GT6):720–724.
Dobry, R., R.S. Ladd, F.Y. Yokel, R.M. Chung, and D. Powell. 1982. Prediction of Pore Water Pressure
Buildup and Liquefaction of Sands during Earthquakes by the Cyclic Strain Method. NBS
Building Science Series 138. Gaithersburg, MD: National Bureau of Standards.
Dobry, R., T. Abdoun, T.D. O’Rourke, and S.H. Goh. 2003. Single pile in lateral spreads: Field bending
moment evaluation. Journal of Geotechnical and Geoenvironmental Engineering 129(10).
Dolbow, J., N. Moes, and T. Belytschko. 2001. An extended finite element method for modeling crack
growth with frictional contact. Computational Methods in Applied Mechanics and Engineering
190(51):6825–6846.
Douglas, J. 2010. Consistency of ground-motion predictions from the past four decades. Bulletin of
Earthquake Engineering 8(6):1515–1526.
Dutton, C.E. 1889. The Charleston earthquake of August 31, 1886. Pp. 203–528 in Ninth Annual Report
of the United States Geological Survey to the Secretary of the Interior 1887-‘88. Washington,
DC. Available at http://pubs.usgs.gov/ar/09/report.pdf; accessed 29 November 2016.
Ebeling, R.M. and E.E. Morrison. 1993. The Seismic Design of Waterfront Retaining Structures.
Technical Report No. R-939. Port Hueneme, CA: Naval Civil Engineering Laboratory. 336 pp.
Elgamal, A., and Z. Yang. 2000. Numerical modeling of liquefaction induced lateral spreading. In
Proceedings of the 12th World Conference on Earthquake Engineering, 30 January–4 February
2000, Auckland, New Zealand. Upper Hutt: New Zealand Society for Earthquake Engineering.
Elgamal, A., M. Zeghal, and E. Parra. 1996. Liquefaction of reclaimed island in Kobe, Japan. Journal of
Geotechnical Engineering 122(1):39–49.
Elgamal, A., Z. Yang, and E. Parra. 2002. Computational modeling of cyclic mobility and post-
liquefaction response. Soil Mechanics and Earthquake Engineering 22:259–271.
Elgamal, A., Z. Yang, E. Parra, and A. Ragheb. 2003. Modeling of cyclic mobility in saturated
cohesionless soils. International Journal of Plasticity 19(6):883–905.
Elgamal, A., J. Lu, and Z. Yang. 2005. Liquefaction induced settlement of shallow foundations and
remediation: 3D numerical simulation. Journal of Earthquake Engineering 9(1):17–45.
Elias, V., J. Welsh, J. Warren, R. Lukas, J.G. Collin, and R.R. Berg. 2006. Ground Improvement
Methods. Reference Manual for NHI Course No. 132034, Report No. FHWA NHI-06-019,
August. Washington, DC: U.S. Federal Highway Administration
Ellison, K. and J.E. Andrade. 2009. Liquefaction mapping in finite element simulations. Journal of
Geotechnical and Geoenvironmental Engineering 135(11):1693–1701.
Faris, A., R.B. Seed, R.E. Kayen, and J. Wu. 2006. A semi-empirical model for the estimation of
maximum horizontal displacement due to liquefaction-induced lateral spreading. In 8th U.S.
National Conference on Earthquake Engineering, 18–22 April, San Francisco, California.
Oakland, CA: Earthquake Engineering Research Institute.
FHWA (Federal Highway Administration). 2006. Seismic Retrofitting Manual for Highway Structures:
Part 1-Bridges. FHWA-HRT-06-032. McLean, VA: U.S. Federal Highway Administration,
Turner-Fairbank Highway Research Center.
Figueroa, J.L., A.S. Saada, and L. Liang. 1995. Effect of the grain size on the energy per unit volume at
the onset of liquefaction. Pp. 197–202 in Third International Conference on Recent Advances in
Geotechnical Earthquake Engineering and Soil Dynamics, Vol. 1, St. Louis, Missouri.
Finn, W. 2003. Landslide-generated tsunamis: Geotechnical considerations. Pure and Applied Geophysics
160:1879–1894.
Finn, W., and D. Liam. 2007. Logical Evaluation of Liquefaction using NBCC 2005 Probabilistic Ground
Accelerations. In Proceedings, 9th Canadian Conference on Earthquake Engineering, 26–29
June, Ottawa.
Finn, W.D.L., S.L. Kramer, T.D. O’Rourke, and T.S. Youd. 2010. Final Report: Technical Issues in
Dispute with EERI MNO-12, Soil Liquefaction During Earthquakes. Report of the Earthquake
REFERENCES 223

Engineering Research Institute Ad Hoc Committee on Soil Liquefaction Earthquakes. Oakland,


CA: Earthquake Engineering Research Institute. Available at
https://www.eeri.org/site/images/email_blasts/Final-Ad-Hoc-Committee-Report-16Aug10.pdf;
accessed 29 November 2016.
Forcellini, D., F.D. Bartola, and A.M. Tarantino. 2013. Liquefaction induced lateral deformations
computational assessment during Tohoku earthquake. ISRN Civil Engineering 2013.
Franke, K.W., and S.L. Kramer. 2014. Procedure for the empirical evaluation of lateral spread
displacement hazard curves. Journal of Geotechnical and Geoenvironmental Engineering
140(1):110–120.
Franke, K.W., and K.M. Rollins. 2013. Simplified hybrid p-y spring model for liquefied soils. Journal of
Geotechnical and Geoenvironmental Engineering 139(4):564–576.
Franke, K.W., A.D. Wright, and C.K. Hatch. 2014. PBliquefY: A new analysis tool for the performance-
based assessment of liquefaction triggering. In Proceedings of the 10th National Conference in
Earthquake Engineering, Earthquake Engineering Research Institute, 21–25 July, Anchorage,
Alaska.
GEER (Geotechnical Extreme Events Reconnaissance). 2010. Geotechnical Reconnaissance of the 2010
Darfield (New Zealand) Earthquake. GEER Association Report No. GEER-024. Geotechnical
Extreme Events Reconnaissance (GEER) Association.
Gelman, A., J.B. Carlin, H.S. Stern, D.B. Dunson, A. Vehtari, and D.B. Rubin. 2014. Bayesian Data
Analysis, 3rd Ed.. Boca Raton FL: CRC Press.
Ghafghazi, M., A. Thurairajah, J.T. DeJong, D.W. Wilson, and R. Armstrong. 2014. Instrumented Becker
Penetration test for improved characterization of gravelly deposits. Geo-Congress 2014 Technical
Papers: Geo-Characterization and Modeling for Sustainability, Geotechnical Special Publication
(GSP) 234:37–46.
Gillette, D.R. 2010. On the Use of Empirical Correlations for Estimating the Residual Undrained Shear
Strength of Liquefied Soils in Dam Foundations. Paper 26 in 5th International Conference on
Recent Advances in Geotechnical Earthquake Engineering and Soil Dynamics, 24 May, San
Diego, California.
Gingery, J.R., A. Elgamal, and J.D. Bray. 2014. Response spectra at liquefaction sites during shallow
crustal earthquakes. Earthquake Spectra 31(4): 2325–2349.
Gohl,W.B., J.A. Howie, and C.E. Rea. 2001. Use of Controlled Detonation of Explosives for Liquefaction
Testing. Paper No. 9.13 in Proceedings, 4th International Conference on Recent Advances in
Geotechnical Earthquake Engineering and Soil Dynamics a Symposium in Honor of Professor W.
D. Liam Finn, 26–31 March, San Diego, California.
Goren, L., E. Aharonov, D. Sparks, and R. Toussaint. 2010. Pore pressure evolution in deforming
granular material: A general formulation and the infinitely stiff approximation. Journal of
Geophysical Research 115(B09216), doi:10.1029/2009JB007191.
Green, R.A. 2001. Energy-based evaluation and remediation of liquefiable soils. Ph.D. diss. Virginia
Polytechnic Institute State University, Blacksburg, VA. 393 pp.
Green, R.A., and J.K. Mitchell. 2003. A Closer Look at Arias Intensity-Based Liquefaction Evaluation
Procedures. Paper Number 94 in Proceedings of the 7th Pacific Conference on Earthquake
Engineering, 13–15 February, University of Canterbury, Christchurch, New Zealand. 9 pp.
Green, R.A., and J.K. Mitchell. 2004. Energy-Based Evaluation and Remediation of Liquefiable Soils.
Pp. 1961–1970 in Geotechnical Engineering for Transportation Projects, Vol. 2, edited by M.
Yegian and E. Kavazanjian. ASCE Geotechnical Special Publication No. 126. Reston, VA:
American Society of Civil Engineers.
Green, R.A., and G.A. Terri. 2005. Number of equivalent cycles concept for liquefaction evaluations—
Revisited. Journal of Geotechnical and Geoenvironmental Engineering 131(4):477–488.
Green, R.A., J.K. Mitchell, and C.P. Polito. 2000. An energy-based pore pressure generation model for
cohesionless soils. In Proceedings: John Booker Memorial Symposium, 16–17 November,
Melbourne, Australia.
224 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Green, R.A., S.F. Obermeier, and S.M. Olson. 2005. Engineering geologic and geotechnical analysis of
paleoseismic shaking using liquefaction effects: Field examples. Engineering Geology 76:263–
293.
Green, R.A., C. Wood, B. Cox, M. Cubrinovski, L. Wotherspoon, B. Bradley, T. Algie, J. Allen, A.
Bradshaw, and G. Rix. 2011. Use of DCP and SASW tests to evaluate liquefaction potential:
Predictions vs. observations during the recent New Zealand earthquakes. Seismological Research
Letters 82(6):927–938.
Green, R.A., M. Cubrinovski, B. Cox, C. Wood, L. Wotherspoon, B. Bradley, and B. Maurer. 2014.
Select liquefaction case histories from the 2010-2011 Canterbury earthquake sequence.
Earthquake Spectra 30:131–153, doi:10.1193/030713EQS066M.
Griffiths, S., and B. Cox. 2012. A comparison of SPT-based empirical liquefaction triggering procedures
for soils at significant depths (+20 m). GeoCongress 2012: 1770-1779.
doi:10.1061/9780784412121.182
Gu, W.H., N.R. Morgenstern, and P.K. Robertson. 1993. Progressive failure of lower San Fernando dam.
Journal of Geotechnical Engineering 119: 333–349.
Gueguen, Ph., M. Langlais, P. Foray, Ch. Rousseau, and J. Maury. 2011. A natural seismic isolating
system: The buried mangrove effects. Bulletin of Seismological Society of America 101:1073–
1080.
Guo, N., and J. Zhao. 2014. A coupled FEM/DEM approach for hierarchical multiscale modeling of
granular media. International Journal for Numerical Methods in Engineering 99(11): 789–818.
Gupta, K.K., and J.L. Meek. 1996. A Brief History of the Beginning of the Finite Element Method.
International Journal for Numerical Methods in Engineering 39:3761–3774.
Gutenberg, B., and C.F. Richter. 1956. Magnitude and Energy of Earthquakes. Annali di Geofisica 9:1–
15.
Gyori, E., L. Toth, Z. Graczer, and T. Katona. 2011. Liquefaction and post liquefaction settlement
assessment – a probabilistic approach. Acta Geodaetica et Geophysica Hungarica 46(3):347–369.
Haase, J.S., Y.S. Choi, and R.L. Nowack. 2011. Liquefaction hazard near the Ohio river from midwestern
scenario earthquakes. Environmental and Engineering Geoscience 17:165–181,
doi:10.2113/gseegeosci.17.2.165.
Hamada, M., S. Yasuda, R. Isoyama, and K. Emoto. 1986. Study on liquefaction induced permanent
ground displacements. Association for the Development of Earthquake Prediction in Japan. 87
pp.
Hamada, M., and T.D. O’Rourke, eds. 1992. Case Studies of Liquefaction and Lifeline Performance
During Past Earthquakes: Volume 1, Japanese Case Studies. Technical Report NCEER-92-0001.
Buffalo, NY: National Center for Earthquake Engineering Research.
Hancock J, Bommer JJ. 2005. The effective number of cycles of earthquake ground motion. Earthquake
Engineering and Structural Dynamics 34(6):637–64.
Harder, L.F., Jr. 1988. Use of penetration tests to determine the cyclic loading resistance of gravelly soils
during earthquake shaking. Ph.D. diss., University of California, Berkeley.
Harder, L.F., Jr., and R.W. Boulanger. 1997. Application of Kσ and Kα correction factors. Pp. 167–190 in
Proceedings of NCEER Workshop on Evaluation of Liquefaction Resistance of Soils, January 5–
6, 1996, Salt Lake City, UT. Buffalo, NY: National Center for Earthquake Engineering Research.
Harder, L.F., Jr., and H.B. Seed. 1986. Determination of penetration resistance for coarse-grained soils
using the Becker hammer drill. Rep. UCB/EERC-86/06. Earthquake Engineering Research
Center, University of California, Berkeley.
Hartford, D.N.D., and G.B. Baecher. 2004. Risk and Uncertainty in Dam Safety. London: Thomas
Telford, Ltd.
Hartvigsen, A.J. 2007. Influence of pore pressures in liquefiable soils on elastic response spectra.
Master’s thesis, University of Washington, Seattle. 150 pp.
REFERENCES 225

Hashash, Y.M.A., D.R. Groholski, and C.A. Phillips. 2010. Recent advances in nonlinear site response
analysis. Paper No. OSP 4 in 5th International Conference on Recent Advances in Geotechnical
Earthquake Engineering and Soil Dynamics, 24 May, San Diego, California.
Hashash, Y.M.A, D.R. Groholski, C.A. Phillips, D. Park, and M. Musgrove. 2012. DEEPSOIL 5.1, User
Manual and Tutorial. 107 pp.
Hashash, Y.M.A., N.A. Abrahamson, S.M. Olson, S. Hague, and B. Kim. 2015. Conditional mean spectra
in site-specific seismic hazard rvaluation for a major river crossing in the central United States.
Earthquake Spectra 31(1):47–69.
Hauksson, E., J. Stock, R. Bilham, M. Boese, X. Chen, E.J. Fielding, J. Galetzka, K.W. Hudnut, K.
Hutton, L.M. Jones, H. Kanamori, P.M. Shearer, J. Steidl, J. Treiman, S. Wei, and W. Yang.
2013. Report of the August 2012 Brawley Earthquake Swarm in Imperial Valley, Southern
California. Seismological Research Letters 84(2):177–189.
Hayati, H., and R.D. Andrus. 2008. Liquefaction potential map of Charleston, South Carolina based on
the 1986 earthquake. Journal of Geotechnical and Geoenvironmental Engineering 134 (6):815–
828, doi:10.1061/(ASCE)1090-0241(2008)134:6(815).
Hayati, H., and R.D. Andrus. 2009. Updated liquefaction resistance correction factors of aged sands.
Journal of Geotechnical and Geoenvironmental Engineering 135(11):1683–1692.
Heidari, T., and R.D. Andrus. 2010. Mapping liquefaction potential of aged soil deposits in Mount
Pleasant, South Carolina. Engineering Geology 112:1–12, doi:10.1016/j.enggeo.2010.02.001.
Heidari, T., and R.D. Andrus. 2012. Liquefaction potential assessment of Pleistocene beach sands near
Charleston, South Carolina. Journal of Geotechnical and Geoenvironmental Engineering
138:1196–1208, doi:10.1061/(ASCE)GT.1943-5606.0000686.
Hill, R. 1958. A general theory of uniqueness and stability in elastic-plastic solids. Journal of the
Mechanics and Physics of Solids 6:236–249.
Hofmann, B.A., D.C. Sego, and P.K. Robertson. 2000. In situ ground freezing to obtain undisturbed
samples of loose sand for liquefaction assessment. Journal of Geotechnical and
Geoenvironmental Engineering 126(11):979–989.
Holzapfel, G. A. 2000. Nonlinear solid mechanics. West Sussex, U.K.: John Wiley & Sons, Inc.
Holzer, T.L., and M.J. Bennett. 2007. Geologic and hydrogeologic controls of boundaries of lateral
spreads: Lessons from USGS case histories. Pp 502–522 in First North American Landslide
Conference. Association of Engineering Geologists Special Publication 23 Available online at
http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.556.4602&rep=rep1&type=pdf;
accessed 29 November 2016.
Holzer, T.L., and T.L. Youd. 2007. Liquefaction, ground oscillation, and soil deformation at the Wildlife
array, California. Bulletin of the Seismological Society of America 97(3):961–976.
Holzer, T., M. Bennett, D. Ponti, and J. Tinsley. 1999. Liquefaction and soil failure during 1994
Northridge earthquake. Journal of Geotechnical and Geoenvironmental Engineering 125:438–
452, doi:10.1061/(ASCE)1090-0241(1999)125:6(438) ER.
Holzer, T., T. Noce, M. Bennett, J. Tinsley, and L. Rosenberg. 2005. Liquefaction at Oceano, California,
during the 2003 San Simeon earthquake. Bulletin of the Seismological Society of America
95:2396–2411, doi:10.1785/0120050078 ER.
Holzer, T.L., M.J. Bennett, T.E. Noce, A.C. Padovani, and J.C. Tinsley III. 2006a. Liquefaction hazard
mapping with LPI in the greater Oakland, California, area. Earthquake Spectra 22(3):693–708.
Holzer, T.L., J.L. Blair, T.E. Noce, M.J. Bennett. 2006b. Predicted liquefaction of east bay fills during a
repeat of the 1906 San Francisco earthquake. Earthquake Spectra 22(S2):S261-277.
Holzer, T.L., T.E. Noce, and M.J. Bennett. 2011. Liquefaction probability curves for surficial geologic
deposits. Environmental and Engineering Geoscience 17:1–21, doi:10.2113/gseegeosci.17.1.1
ER.
Huang, Y. 2008. Performance-based design and evaluation for liquefaction-related seismic hazard. Ph.D.
diss., University of Washington, Seattle.
226 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Huang, A.B., Y.Y. Tai, W.F. Lee, and K. Ishihara. 2008. Sampling and field characterization of the silty
sand in Central and Southern Taiwan. Pp. 1457–1463 in 3rd International Conference on Site
Characterization (ISC-3). Taipei, Taiwan: Taylor & Francis.
Hughes, T.J.R. 1987. The Finite Element Method: Linear Static and Dynamic Finite Element Analysis.
Englewood Hills, NJ: Prentice Hall.
Hurst, A., A. Scott, and M. Vigorito. 2011. Physical characteristics of sand injectites. Earth-Science
Reviews 106:215–246, doi:10.1016/j.earscirev.2011.02.004.
Hushmand, B., R.F. Scott, and C.B. Crouse. 1991. In situ calibration of USGS piezometer installations.
Pp. 49–61 in Recent Advances in Instrumentation, Data Acquisition, and Testing in Soil
Dynamics, edited by S. K. Bhatia and G. W. Blaney. ASCE Special Publication 29. Reston, VA:
American Society of Civil Engineers.
Hushmand, B., R.F. Scott, and C.B. Crouse. 1992. In-place calibration of USGS pore pressure transducers
at Wildlife Liquefaction Site, California, USA. Pp. 1263–1268 in Proceedings of the 10th
Earthquake Engineering World Conference. Rotterdam, the Netherlands: Balkema.
Hynes, M.E., and R.S. Olsen. 1999. Influence of confining stress on liquefaction resistance. Pp. 145–152
in Proceedings International Symposium on the Physics and Mechanics of Liquefaction.
Rotterdam, the Netherlands: Balkema.
Iai, S. 1998. Rigid and flexible walls during the Kobe earthquake. Pp. 108–127, Paper No. SOA-4 in
Proceedings of the Fourth International Conference on Case Histories in Geotechnical
Engineering. Rolla: Missouri University of Science and Technology.
Iai, S., H. Tsuchida, and K. Koizumi. 1986. A new criterion for assessing liquefaction potential using
grain size accumulation curve and N-value. Report of the Port and Harbour Research Institute 25
(3):125–234 (in Japanese).
Iai, S., H. Tsuchida, and K. Koizumi. 1989. A liquefaction criterion based on field performances around
seismograph stations. Soils and Foundations 29(2):52–68.
Iai, S., Y. Matsunaga, and T. Kameoka. 1990. Strain space plasticity model for cyclic mobility. Report of
the Port and Harbour Research Institute 29:27–56.
Iai, S., Y. Matsunaga, and T. Kameoka. 1992. Strain space plasticity model for cyclic mobility. Soil and
Foundations 32(2):1–15.
Iai, S., T. Tobita, and Y. Tamari. 2008. Seismic performance and design of port structures. Pp. 1–16 in
Proceedings, Geotechnical Earthquake Engineering and Soil Dynamics IV, ASCE, Sacramento,
California. Reston, VA: American Society of Civil Engineers.
Iai, S., T. Tobita, and O. Ozutsumi. 2011. Induced fabric under cyclic and rotational loads in a strain
space multiple mechanism model for granular materials. International Journal for Numerical and
Analytical Methods in Geomechanics 37:1326–1336.
ICC (International Code Council). 2011. 2012 International Building Code. Country Club Hills, IL:
International Code Council.
ICOLD (International Commission on Large Dams). 2002. Earthquake design and evaluation of
structures appurtenant to dams. Bulletin 123. Paris, France: Committee on Seismic Aspects of
Dam Design, International Commission on Large Dams.
Idriss, I.M. 1985. Evaluating seismic risk in engineering practice. Pp. 255–320 in Proceedings of the
Eleventh International Conference on Soil Mechanics and Foundation Engineering, Eleventh
International Conference on Soil Mechanics and Foundation Engineering, Vol. 1, 12–16 August,
San Francisco, Califormia. Boston, MA: A.A. Balkema.
Idriss, I.M. 1999. An update to the Seed-Idriss simplified procedure for evaluating liquefaction potential.
In Proceedings of the TRB Workshop on New Approaches to Liquefaction Analysis. Special
Report Number MCEER-99-SP04. Washington, DC: Federal Highway Administration.
Idriss, I.M., and R.W. Boulanger. 2007a. Residual shear strength of liquefied soils. Pp. 621–634 in
Proceedings of the Modernization and Optimization of Existing Dams and Reservoirs: 27th
Annual USSD Conference, Philadelphia, Pennsylvania, March 5-9 2007. Denver, CO: United
States Society on Dams.
REFERENCES 227

Idriss, I.M., and R.W. Boulanger. 2007b. SPT- and CPT-based relationships for the residual shear
strength of liquefied soils. Pp. 1–22 in Earthquake Geotechnical Engineering, 4th International
Conference on Earthquake Geotechnical Engineering – Invited Lectures, edited by K.D. Pitilakis.
The Netherlands: Springer.
Idriss, I.M., and R.W. Boulanger. 2008. Soil Liquefaction During Earthquakes. Earthquake Engineering
Research Institute MNO 12. Oakland, CA: Earthquake Engineering Research Institute.
Idriss, I.M., and R.W. Boulanger. 2010. SPT-Based Liquefaction Triggering Procedures. Report No.
UCD/CGM-10-02. Center for Geotechnical Modeling, Department of Civil and Environmental
Engineering, University of California, Davis. 259 pp. Available at
http://faculty.engineering.ucdavis.edu/boulanger/wp-
content/uploads/sites/71/2014/09/Idriss_Boulanger_SPT_Liquefaction_CGM-10-02.pdf; accessed
29 November 2016.
Idriss, I.M. and R.W. Boulanger. 2012. Examination of SPT-based liquefaction triggering correlations.
Earthquake Spectra 28(3):989–1018.
International Navigation Association. 2001. Seismic Design Guidelines for Port Structures. Working
Group No. 34 of the Maritime Navigation Commission. Lisse, the Netherlands: A.A. Balkema.
Ishihara, K. 1984. Post-earthquake failure of a tailings dam due to liquefaction of the pond deposit. Pp.
1129–1143 in Proceedings, International Conference on Case Histories in Geotechnical
Engineering, 6–11 May, Rolla, Missouri.
Ishihara, K. 1985. Stability of natural deposits during earthquakes. Pp. 321-376 in Proceedings of the
11th International Conference on Soil Mechanics and Foundation Engineering, San Francisco,
California.
Ishihara, K. 1993. Liquefaction and flow failure during earthquakes. Géotechnique 43(3):351–415.
Ishihara, K. 1996. Soil Behaviour in Earthquake Geotechnics. Oxford, U.K.: Oxford Science
Publications, Clarendon Press.
Ishihara, K., and M. Cubrinovski. 2004. Performance of piles as evaluated by three-layer model. Pp. 167–
191 in Proceedings of Geotrans 2004: Geotechnical Engineering for Transportation Projects.
ASCE Geotechnical Special Publication 126. Reston, VA: American Society of Civil Engineers,
doi:10.1061/40744(154)5.
Ishihara, K. and M. Yoshimine. 1992. Evaluation of settlements in sand deposits following liquefaction
during earthquakes. Soils and Foundations 32(1):173–188.
Ishihara, K., R. Tatsuoka, and S. Yasuda. 1975. Undrained deformation and liquefaction of sand under
cyclic stresses. Soils and Foundations 15(1):29–44.
Iverson, R.M., and D.L. George. 2014. A depth-averaged debris-flow model that includes the effects of
evolving dilatancy. I. Physical basis. Proceedings of the Royal Society A 470:20130819.
Iwasaki, T., F. Tatsuoka, K. Tokida, and S. Yasuda. 1978. A practical method for assessing soil
liquefaction potential based on case studies at various sites in Japan. Pp. 885–896 in Proceedings
of the 2nd International Conference on Microzonation, San Francisco, California.
Jaeger, H.M., and S.R. Nagel. 1992. Physics of the granular state. Science 255(5051):1523–1531.
Janecke, S.U., R.Q. Oaks, A.J. Knight, D. Nutt, and T.M. Rittenour. 2013. Large liquefaction features and
evidence for earthquakes induced by Lake Bonneville in Cache Valley; A progress report
[abstract]. Abstracts with Programs - Geological Society of America Annual Meeting, Denver,
Colorado 45(7):711.
Jefferies, M.G. and K. Been. 2006. Soil Liquefaction: A Critical State Approach. London: Taylor &
Francis.
Jibson, R.W. 2007. Regression models for estimating coseismic landslide displacement. Engineering
Geology 91(2):209–218.
Juang, C. H., T. Jiang, and R.D. Andrus. 2002. Assessing probability-based methods for liquefaction
potential evaluation. Journal of Geotechnical and Geoenvironmental Engineering 128(7):580–
589.
228 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Juang, C.H., C.-N. Liu, C.-H. Chen, J.-H. Hwang, and C.-C. Lu. 2008. Calibration of liquefaction
potential index: A re-visit focusing on a new CPTU model. Engineering Geology 102(1-2):19–30.
Juang, C.-H., D.K. Li, S.Y. Fang, Z. Liu, and H. Khor. 2008. Simplified procedure for developing joint
distribution of amax and Mw for probabilistic liquefaction hazard analysis. Journal of Geotechnical
and Geoenvironmental Engineering 134(8):1050–1058.
Kammerer, A. 2002. Undrained response of Monterey 0/30 sand under multidirectional cyclic simple
shear loading conditions. Ph.D. diss., University of California, Berkeley.
Kavazanjian, et al. 1998. Geotechnical Earthquake Engineering: Reference Manual, edited by G.
Munfakh, E. Kavazanjian, Jr., N. Matasovic, T. Hadj-Hamou, and J.-N. Wang. Training Course
in Geotechnical and Foundation Engineering, NHI Course NO. 13239 – Module 9. Publication
No. FHWA HI-99-012. Arlington, VA: Federal Highway Administration, National Highway
Institute, U.S. Department of Transportation. Available at
https://ntrl.ntis.gov/NTRL/dashboard/searchResults/titleDetail/PB99142531.xhtml#; accessed 18
November 2016.
Kavazanjian Jr., E., J.-N.J. Wang, G.R. Martin, A. Shamsabadi, I. Lam, S.E. Dickenson, C.J. Hung, 2011.
LRFD Seismic Analysis and Design of Transportation Geotechnical Features and Structural
Foundations Reference Manual. NHI Course No. 130094 Reference Manual. Geotehnical
Engineering Circular No. 3. Washington, DC: National Highway Institute, U.S. Department of
Transportation, Federal Highway Administration. Available at
https://www.fhwa.dot.gov/engineering/geotech/pubs/nhi11032/nhi11032.pdf; accessed
November 29, 2016.
Kayen, R., and J.K. Mitchell. 1997. Assessment of liquefaction potential during earthquakes by Arias
Intensity. Journal of Geotechnical and Geoenvironmental Engineering 123(12):1162–1174.
Kayen, R.R., R.E.S. Moss, E.R. Thompson, R.B. Seed, K.O. Cetin, A. Derkiureghian, Y. Tanaka, and K.
Tokimatsu. 2013. Shear wave velocity-based probabilistic and deterministic assessment of
seismic soil liquefaction potential. Journal of Geotechnical and Geoenvironmental Engineering
139(3):407–419.
Kishida, T., and C.-C. Tsai. 2014. Seismic demand of the liquefaction potential with equivalent number
of cycles for probabilistic seismic hazard analysis. Journal of Geotechnical and
Geoenvironmental Engineering 140(3), doi:10.1061/(ASCE)GT.1943-5606.0001033.
Kishida, T., R.W. Boulanger, N.A. Abrahamson, M.W. Driller, and T. M. Wehling. 2009. Seismic
response of levees in Sacramento-San Joaquin Delta. Earthquake Spectra 25(3):557-582.
Klecka, W.R. 1980. Discriminant Analysis, QASS No. 19. Newbury Park, CA: Sage Publications.
Koester, J.P. 1992. The influence of test procedure on correlation of Atterberg limits with liquefaction in
fine-grained soils. Geotechnical Testing Journal. 15(4): 352–360.
Kokusho, T. 2007. Liquefaction strengths of poorly-graded and well-graded granular soils investigated by
lab tests. Pp. 159–184 in 4th International Conference on Earthquake Geotechnical Engineering,
Thessaloniki, Greece. The Netherlands: Springer.
Kokusho, T. 2013. Liquefaction potential evaluations: Energy-based method versus stress-based method.
Canadian Geotechnical Journal 50(10):1088–1099.
Kottke, A. R., and E.M. Rathje. 2008. Technical manual for Strata. University of California, Berkeley.
Kovacs, W.D., L.A. Salomone, and F.Y. Yokel. 1981. Energy measurement in the standard penetration
test. Washington, DC: U.S. Deptartment of Commerce, National Bureau of Standards.
Kramer, S.L. 1996. Geotechnical Earthquake Engineering. Upper Saddle River, NJ: Prentice Hall.
Kramer, S.L. 2008. Evaluation of liquefaction hazards in Washington State. Report WA-RD 668.1.
Seattle: Washington State Transportation Center. 152 pp. Available at
http://www.digitalarchives.wa.gov/Record/View/AB470112278FCFE11C7913CC84026574;
accessed November 29,2016.
Kramer, S.L., 2011. Performance-based design in geotechnical earthquake engineering practice. Invited
State-of-the-Art Paper, Proceedings, 5th International Conference on Earthquake Geotechnical
Engineering, Santiago, Chile, 34 pp.
REFERENCES 229

Kramer, S.L. 2013. Performance-based design methodologies for geotechnical earthquake engineering.
Bulletin of Earthquake Engineering 12(3): 1049-1070.
Kramer, S.L., and P. Arduino. 1999. Constitutive modeling of cyclic mobility and implications for site
response. Pp. 1029–1034 in Proceedings of Second International Conference on Earthquake
Geotechnical Engineering, Lisbon, Portugal.
Kramer, S.L., and A.W.M. Elgamal. 2001. Modeling soil liquefaction hazards for performance-based
earthquake engineering. Pacific Earthquake Engineering Research Center (PEER) Report
2001/13. University of California, Berkeley.
Kramer, S.L., and Y.-M. Huang. 2010. Performance-based assessment of liquefaction hazards. In
Proceedings of the 9th U.S. National and 10th Canadian Conference on Earthquake Engineering,
Toronto, Canada. 10 pp.
Kramer, S.L., and R.T. Mayfield. 2005. Performance-based liquefaction hazard evaluation. In
Proceedings, GeoFrontiers 2005: ASCE, Austin, Texas. Reston, VA: American Society of Civil
Engineers.18 pp.
Kramer, S.L., and R.T. Mayfield. 2007. The return period of liquefaction. Journal of Geotechnical and
Geoenvironmental Engineering 133(7):802–813.
Kramer, S.L., and J.P. Stewart. (in preparation). Geotechnical Earthquake Engineering, 2nd Ed. Prentice
Hall.
Kramer, S.L, and C. Wang. 2015. Empirical model for estimation of the residual strength of liquefied soil.
Journal of Geotechnical and Geoenvironmental Engineering 141(9):04015038.
Kramer, S.L., R.T. Mayfield, and D.G. Anderson. 2006. Performance-based liquefaction hazard
evaluation: Implications for codes and standards. Paper No. 888 in Proceedings, Eighth U.S.
National Conference on Earthquake Engineering, San Francisco, California.
Kramer, S.L., P. Arduino, and H.S. Shin. 2009. Development of performance criteria for foundations and
earth structures. Pp. 107–120 in Performance-Based Design in Earthquake Geotechnical
Engineering: From Case History to Practice, edited by T. Kokusho, Y. Tsukamoto, and M.
Yoshimine, London, U.K.: Taylor & Francis.
Kramer, S.L., A.J. Hartvigsen, S.S. Sideras, and P.T. Ozener. 2011. Site response modeling in liquefiable
soil deposits. In Proceedings of Fourth IASPEI / IAEE International Symposium: Effects of
Surface Geology on Seismic Motion. University of California, Santa Barbara.
Kramer, S.L., B. Astaneh, P. Ozener, and S. Sideras. 2013. Effects of liquefaction on ground surface
motions. In Proceedings, International Conference on Earthquake Geotechnical Engineering,
ICEGE Istanbul 2013 – From Case History to Practice, Istanbul, Turkey.
Kutter, B.L. 1992. Dynamic centrifuge modeling of geotechnical structures. Transportation Research
Record: Journal of the Transportation Research Board 1336: 24–30.
Kwasinski, A., J. Eidinger, A. Tang, and C. Tudo-Bornarel. 2014. Performance of electric power systems
in the 2010-2011 Christchurch, New Zealand, earthquake sequence. Earthquake Spectra
30(1):205–230.
Lacy, S.J., and J.H. Prevost. 1987. Nonlinear seismic response analysis of earth dams. Soil Dynamics and
Earthquake Engineering 6(1):48–63.
Ladd, R.S. 1974. Specimen preparation and liquefaction of sands. Journal of the Geotechnical
Engineering Division 100(GT10):1180–1184.
Law, K.T., Y.L. Cao, and G.N. He. 1990. An energy approach for assessing seismic liquefaction
potential. Canadian Geotechnical Journal 27(3):320–329, 10.1139/t90-043.
Ledezma, C., and J.D. Bray. 2010. Probabilistic performance-based procedure to evaluate pile
foundations at sites with liquefaction-induced lateral displacement. Journal of Geotechnical and
Geoenvironmental Engineering 136(3):464–476.
Ledezma, C., T. Hutchinson, S.A. Ashford, R. Moss, P. Arduino, J.D. Bray, S. Olson, Y.M.A. Hashash,
R. Verdugo, D. Frost, R. Kayen, and K. Rollins. 2012. Effects of ground failure on bridges, roads,
and railroads. Earthquake Spectra 28(S1):S119–S143.
230 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Lee, D.H., C.S. Ku, and H. Yuan. 2003. A study of liquefaction risk potential at Yuanlin, Taiwan.
Engineering Geology 71: 97–117.
Lee, J., and R.A. Green. 2008. An empirical earthquake Arias Intensity relationship for the Western US.
Paper 180 in Proceedings of the Workshop on Signal Processing and Its Applications (WoSPA
2008), 18–20 March, University of Sharjah, United Arab Emirates.
Lee, J., and R.A. Green. 2010. An empirical Arias Intensity relationship for rock sites in stable
continental regions. In Proceedings of the 7th International Conference on Urban Earthquake
Engineering & 5th International Conference on Earthquake Engineering, 3–5 March, Tokyo,
Japan.
Lee, J., and R.A. Green. 2015. Empirical predictive relationship for seismic lateral displacement of
slopes. Geotechnique 65(5):374–390.
Lee, J., R.A. Green, and R. Finch. 2010. An Empirical Predictive Relationship for Assessing the Seismic
Stability of Slopes. In Proceedings of the Fifth International Conference on Recent Advances in
Geotechnical Earthquake Engineering and Soil Dynamics and Symposium in Honor of Professor
I.M. Idriss, 24–29 May, San Diego, California.
Lenz, A., and L.G. Baise. 2007. Spatial variability of liquefaction potential in regional mapping using
CPT and SPT data. Soil Dynamics and Earthquake Engineering 27:690–702.
Leprince, S., S. Barbot, F. Ayoub, and J. Avouac. 2007. Automatic and precise orthorectification,
coregistration, and subpixel correlation of satellite images, application to ground deformation
measurements. IEEE Transactions on Geoscience and Remote Sensing 45(6):1529–1558.
Li, B., F. Habbal, and M. Ortiz. 2010. Optimal transportation meshfree approximation schemes for fluid
and plastic flows. International Journal for Numerical Methods in Engineering 83(12):1541–
1579.
Li, X.S., H.Y. Ming, and Z.Y. Cai. 2000. Constitutive modeling of flow liquefaction and cyclic mobility.
Pp. 81–98 in Geotechnical Special Publication 110, Geo-Denver 2000: Computer Simulation of
Earthquake Effects, edited by K. Arulanandan, A. Anadarajah, and X.S. Li. 5–8 August 5-8,
Denver, Colorado. Reston, VA: American Society of Civil Engineers.
Liang, L. 1995. Development of an Energy Method for Evaluating the Liquefaction Potential of a Soil
Deposit. Ph.D. diss., Case Western Reserve University, Cleveland, Ohio. 281pp.
Liao, S.S.C., and R.V. Whitman. 1986. A Catalog of Liquefaction and Non-Liquefaction Occurrences
during Earthquakes. Research Report. Cambridge: Department of Civil Engineering,
Massachusetts Institute of Technology.
Liao, S.S.C., D. Veneziano, and R.V. Whitman. 1988. Regression models for evaluating liquefaction
probability. Journal on Geotechnical Engineering Division 114(4):389–411.
Lienkaemper, J.J., and P.L. Williams. 2007. A record of large earthquakes on the southern Hayward Fault
for the past 1800 years. Bulletin of the Seismological Society of America 97:1803–1819,
doi:10.1785/0120060258.
Lim, K.W., K. Krabbenhoft, and J.E. Andrade. 2014. On the contact treatment of non-convex particles in
the granular element method. Computational Particle Mechanics 1:257–275.
Liu, A.H., J.P. Stewart, N.A. Abrahamson, and Y. Moriwaki. 2001. Equivalent number of uniform stress
cycles for soil liquefaction analysis. Journal of Geotechnical and Geoenvironmental Engineering
127(12):1017–1026.
Liu, L., and R. Dobry. 1995. Effect of liquefaction on lateral response of piles by centrifuge model tests.
National Center for Earthquake Engineering Research (NCEER) Bulletin Annual 9(1):7–11.
Liu, W.K., E.G. Karpov, and H.S. Park. 2006. Nano Mechanics and Materials. Chichester, U.K.: John
Wiley & Sons, Inc.
Lowe, D.R. 1975. Water escape structures in coarse-grained sediments. Sedimentology 22:157–204.
Lu, J., J. Peng, A. Elgamal, Z. Yang, and K.H. Law. 2004. Parallel finite element modeling of earthquake
ground response and liquefaction. Earthquake Engineering and Engineering Vibration 3(1):23–
37.
Lubliner, J. 1990. Plasticity Theory. New York: Macmillan Publishing Company.
REFERENCES 231

Manzari, M.T. 1996. Seismic analysis of soil embankments. In 11 WCEE: Eleventh World Conference on
Earthquake Engineering, 23–28 June, Acapulco, Mexico. Oxford, U.K.: Elsevier Science Ltd.
Manzari, M.T., and Y.F. Dafalias. 1997. A critical state two-surface plasticity model for sands.
Géotechnique 47(2):255–272.
Manzari, M.T., B.L. Kutter, M. Zeghal, S. Iai, T. Tobita, S.P.G. Madabhushi, S.K. Haigh, L. Mejia, D.A.
Gutierrez, R.J. Armstrong, M.K. Sharp, Y.M. Chen, and Y.G. Zhou. 2014. LEAP projects:
concept and challenges. In Geotechnics for Catastrophic Flooding Events, edited by S. Iai.
London: CRC Press.
Marcuson III, W.F., R.F. Ballard, Jr., and R.H. Ledbetter. 1979. Liquefaction failure of tailings dams
resulting from the Near Izu Oshima earthquake, 14 and 15 January, 1978. Proceedings of the 6th
Pan-American Conference of Soil Mechanics and Foundation Engineering, Lima, Peru.
Marcuson III, W.F., and Bieganousky, W.A. 1977. Laboratory standard penetration tests on fine sands.
Journal of the Geotechnical Engineering Division, ASCE 103(GT6):565-588.
Marcuson, W.F., M. E. Hynes, and A.G. Franklin. 2007. Seismic design and analysis of embankment
dams: The state of practice. In Proceedings of Fourth Civil Engineering Conference in the Asian
Region, Taipei, Taiwan.
Marrone, J., F. Ostadan, R. Youngs, and J. Itemizer. 2003. Probabilistic liquefaction hazard evaluation:
Method and application. Paper No. M02-1 in Proceedings of the 17th International Conference
Structural Mechanics in Reactor Technology (Smart 17), Prague, Czech Republic.
Martin, G.R., W.D.L. Finn, and H.B. Seed. 1975. Fundamentals of liquefaction under cyclic loading.
Journal of the Geotechnical Engineering Division 101(GT5):423–438.
Martin, G.R., M.L. Marsh, D.G. Anderson, R.L. Mayes, and M.S. Power. 2002. Recommended design
approach for liquefaction induced lateral spreads. In Proceedings of the 3rd National Seismic
Conference and Workshop on Bridges and Highways, 28 April–1 May, Portland, Oregon.
Washington, DC: Federal Highway Administration.
Martin, J.G., and E.M. Rathje. 2014. Lateral spread deformations from the 2010-2011 New Zealand
earthquakes measured from satellite images and optical image correlation. In Proceedings of the
10th U.S. National Conference in Earthquake Engineering, 21–25 July, Anchorage, Alaska.
Oakland, CA: Earthquake Engineering Research Institute.
Martin, J.R., and G.W. Clough. 1994. Seismic parameters from liquefaction evidence. Journal of
Geotechnical Engineering 120(8):1345–1361.
Martin, M.E., and J. Bourgeois. 2012. Vented sediments and tsunami deposits in the Puget Lowland,
Washington—differentiating sedimentary processes. Sedimentology 59:419–444,
doi:10.1111/j.1365-3091.2011.01259.x.
Mast, C.M. 2013. Modeling landslide induced flow interactions with structures using the material point
method. Ph.D. diss., University of Washington, Seattle.
Matasovic, N., and G. Ordonez. 2011. DMOD 2000: A Computer Program for Nonlinear Seismic
Response Analysis of Horizontally Layered Soil Deposits, Earthfill Dams and Solid Waste
Landfills. Lacey, WA: GeoMotions, LLC.
Matasovic, N., and Y. Hashash. 2012. Practices and Procedures for Site-Specific Evaluations of
Earthquake Ground Motions: A Synthesis of Highway Practice. National Cooperative Highway
Research Project Synthesis Report No. 428. Washington, DC: Transportation Research Board, 79
pp.
Matasovic, N., and M. Vucetic. 1993. Cyclic Characterization of Liquefiable Sands. Journal of
Geotechnical and Geoenvironmental Engineering 119(11):1805–1822.
Mathias, S.A., A.P. Butler, and H. Zhan. 2008. Approximate solutions for Forchheimer flow to a well.
Journal of Hydraulic Engineering 134(9):1318–1325.
Maurer, B.W., R.A. Green, M. Cubrinovski, and B.A. Bradley. 2014a. Assessment of aging correction
factors for liquefaction resistance at sites of recurrent lLiquefaction. In Proceedings of the 10th
U.S. National Conference on Earthquake Engineering, 21–25 July, Anchorage, Alaska. Oakland,
CA: Earthquake Engineering Research Institute. 11 pp.
232 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Maurer, B.W., R.A. Green, M. Cubrinovski, and B.A. Bradley. 2014b. Evaluation of the liquefaction
potential index for assessing liquefaction hazard in Christchurch, New Zealand. Journal of
Geotechnical and Geoenvironmental Engineering 140(7), doi:10.1061/(ASCE)GT.1943-
5606.0001117.
Maurer, B.W., R.A. Green, M. Cubrinovski, and B.A. Bradley. 2015a. Fines-content effects on
liquefaction hazard evaluation for infrastructure in Christchurch, New Zealand. Soil Dynamics
and Earthquake Engineering 76:58–68.
Maurer, B.W., R.A. Green, M. Cubrinovski, and B.A. Bradley. 2015b. Assessment of CPT-Based
Methods for Liquefaction Evaluation in a Liquefaction Potential Index (LPI) Framework.
Geotechnique 65(5):328–336.
Maurer, B.W., R.A. Green, and O.-D.S. Taylor. 2015c. Moving towards an improved index for assessing
liquefaction hazard: Lessons from historical data. Soils and Foundations 55(4):778–787.
Mayfield, R., S. Kramer, and Y. Huang. 2010. Simplified approximation procedure for performance-
based evaluation of liquefaction potential. Journal of Geotechnical and Geoenvironmental
Engineering 136(1):140–150.
Mayne, P.W. 2007. Cone Penetration Testing. NCHRP Synthesis 368. Washington, DC: Transportation
Research Board. 126 pp.
McCulloch, D.S. and M.G. Bonilla. 1970. Effects of the Earthquake of March 27, 1964, on The Alaska
Railroad. Geological Survey Professional Paper 545-D. Washington, DC: U.S. Department of the
Interior, U.S. Government Printing Office.
McGann, C.R., and P. Arduino. 2014. Numerical assessment of three dimensional foundation pinning
effects during lateral spreading at the Mataquito River Bridge. Journal of Geotechnical and
Geoenvironmental Engineering 140(8).
McGann, C.R., P. Arduino, and P. Mackenzie-Helnwein. 2012. Simplified procedure to account for a
weaker soil layer in lateral load analysis of single piles. Journal of Geotechnical and
Geoenvironmental Engineering 138(9):1129–1137.
Mejia, L.H. 1998. Liquefaction at Moss Landing. In The Loma Prieta, California, earthquake of October
17, 1989—Liquefaction, edited by T.L. Holzer. B129–B150. U.S. Geological Survey Professional
Paper 1551B. Denver, CO: U.S. Geological Survey.
Mejia, L.H., and Dawson, E.M. 2006. Earthquake deconvolution for FLAC. In Proceedings, Fourth
International FLAC Symposium on Numerical Modeling in Geomechanics, 29–31 May, Madrid,
Spain.
Miner, M.A. 1945. Cumulative damage in fatigue. Journal of Applied Mechanics. 12(3):A159–A164.
Mitchell, J.K. 1986. Practical problems from surprising soil behavior. Journal of Geotechnical
Engineering 112(3):259–289.
Mitchell, J.K. 2008. Aging of sand –– A continuing enigma? In Proceedings, 6th International
Conference on Case Histories in Geotechnical Engineering, Arlington, Virginia.
Mitchell, J.K. and K. Soga. 2005. Fundamentals of Soil Behavior, 3rd Ed. Hoboken, NJ: John Wiley &
Sons, Inc. 577 pp.
Mitchell, J.K., and Z.V. Solymar. 1984. Time-dependent strength gain in freshly deposited or densified
sand. Journal of Geotechnical Engineering 110(11):1559–1576.
Moin, P. 2001. Fundamentals of engineering numerical analysis. Cambridge, U.K.: Cambridge
University Press.
Monaco, P., S. Marchetti, G. Totani, and M. Calabrese. 2005. Sand liquefiability assessment by flat
dilatometer test (DMT). Pp. 2693–2697 in Proceedings of the XIV European Conference on Soil
Mechanics and Geotechnical Engineering, 12–16 September, Osaka, Japan.
Montgomery, J., R.W. Boulanger, and L.F. Harder, Jr. 2012. Examination of the Ks overburden correction
factor on liquefaction resistance. Report No. UCD/CGM-12-02. Center for Geotechnical
Modeling, Department of Civil and Environmental Engineering, University of California, Davis.
44 pp.
REFERENCES 233

Moss, R.E.S. 2003. CPT-based probabilistic assessment of seismic soil liquefaction initiation. Ph.D. diss.,
University of California, Berkeley.
http://digitalcommons.calpoly.edu/cgi/viewcontent.cgi?article=1039&amp;context=cenv_fac;
accessed 29 November 2016.
Moss, R.E.S., and G. Chen. 2008. Comparing liquefaction procedures in the U.S. and China. In 14th
World Conference on Earthquake Engineering, 12–17 October, Beijing, China.
Moss, R.E.S., R.B. Seed, R.E. Kayen, J.P. Stewart, A. Der Kiureghian, and K.O. Cetin. 2006. CPT-based
probabilistic and deterministic assessment of in situ seismic soil liquefaction potential. Journal of
Geotechnical and Geoenvironmental Engineering 132(8):1032–1051.
Motamed, R., and Towhata, I. 2010. Mitigation measures for pile groups behind quay wall models
undergoing lateral flow of liquefied soil: Shaking table results. Soil Dynamics and Earthquake
Engineering 30(10):1043–1060.
Muir Wood, D. 1990. Soil Behaviour and Critical State Soil Mechanics. Cambridge, U.K.: Cambridge
University Press.
Mulilis, J.P., H.B. Seed, C.K. Chan, J.K. Mitchell, and K. Arulanandan. 1977. Effects of sample
preparation on sand liquefaction. Journal of the Geotechnical Engineering Division
103(GT2):91–108.
Mulilis, J.P., C.K. Chan, and B.H. Seed. 1975. The Effects of Method of Sample Preparation on the
Cyclic Stress-Strain Behavior of Sands. Report Number UCB/EERC-75/18. Earthquake
Engineering Research Center, University of California, Berkeley. 138 pp.
Mulilus, J.P., K. Arulanandan, J.K. Mitchell, C.K. Chan, and H.B. Seed. 1977. Effects of sample
preparation on sand liquefaction. Journal of the Geotechnical Engineering Division, ASCE
103(GT2):99–108.
Nakata, T., and K. Shimazaki. 1997. Geo-slicer, a newly invented soil sampler for high-resolution active
fault studies. Journal of Geography 106:59–69.
Nakata, T., and K. Shimazaki. 2000. Geo-slicer, a new soil sampler for paleoseismological studies. Pp.
315–318 in Proceedings of the Hokudan International Symposium on Active Faulting.
NCHRP (National Cooperative Highway Research Program). 2002. Comprehensive Specifications for the
Seismic Design of Bridges. Report 472. Washington, DC: Transportation Research Board.
NEHRP (National Earthquake Hazards Reduction Program). 2006. Next-Generation Performance-Based
Seismic Design Guidelines: Program plan for new and existing buildings. Prepared by Applied
Technology Council, Redwood, California. FEMA 445. Available online at
http://www.nehrp.gov/pdf/fema445.pdf; accessed 26 October 2016.
Nemat-Nasser, S., and A. Shokooh. 1977. A Unified Approach to Densification and Liquefaction of
Cohesionless Sand. Technical Report No. 77-10-3. Evanston, IL: Earthquake Research and
Engineering Laboratory, Department of Civil Engineering. 44 pp.
Nemat-Nasser, S., and A. Shokooh. 1979. A unified approach to densification and liquefaction of
cohesionless sand in cyclic shearing. Canadian Geotechnical Journal 16:659–678.
Newmark, N.M. 1965. Effects of earthquakes on dams and embankments. Géotechnique 15(2):139–160.
Nicot, F., F. Darve, and RNVO Group. 2005. A multi-scale approach to granular materials. Mechanics of
Materials 37(9):980–1006.
Nigbor, R.L. and T. Imai. 1994. The suspension P-S velocity logging method. Pp. 57-61 in Geophysical
Characterization of Sites, edited by R.D. Woods. International Society for Soil Mechanics and
Geotechnical Engineering Special Publication TC 10. New York: International Science Publisher.
Nitka, M., G. Combe, C. Dascalu, and J. Desrues. 2011. Two-scale modeling of granular materials: A
DEM-FEM approach. Granular Matter 13(3):277–281.
Nova, R. 1994. Controllability of the incremental response in soil specimens subjected to arbitrary
loading programmes. Journal of the Mechanical Behavior of Materials 5(2):193–202.
NRC (National Research Council). 1985. Liquefaction of Soils During Earthquakes. Committee on
Earthquake Engineering, Commission on Engineering and Technical Systems. Washington, DC:
National Academy Press.
234 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Obermeier, S.F. 1996. Use of liquefaction-induced features for paleoseismic analysis — An overview of
how seismic liquefaction features can be distinguished from other features and how their regional
distribution and properties of source sediment can be used to infer the location and strength of
Holocene paleo-earthquakes. Engineering Geology 44:1–76, doi:10.1016/S0013-7952(96)00040-
3.
Obermeier, S.F., and Dickenson, S.E. 2000. Liquefaction evidence for the strength of ground motions
resulting from late Holocene Cascadia subduction earthquakes, with emphasis on the event of
1700 A.D. Bulletin of the Seismological Society of America 90:876–896.
Obermeier, S.F., N.R. Bleuer, C.A. Munson, P.J. Munson, W.S. Martin, K.M. McWilliams, D.A.
Tabaczynski, J.K. Odum, M. Rubin, and D.L. Eggert. 1991. Evidence of strong earthquake
shaking in the lower Wabash Valley from prehistoric liquefaction features. Science 251:1061–
1063.
Ohsaki, Y. 1966. Niigata Earthquakes, 1964, Building Damage and Soil Condition. Soils and
Foundations 6(2):14–37.
Okamura, M., and Y. Soga. 2006. Effects of pore fluid compressibility on liquefaction resistance of
partially saturated sand. Soils and Foundations 46(5):695–700.
Olson, S.M., and C.I. Johnson. 2008. Analyzing liquefaction-induced lateral spreads using strength ratios.
Journal of Geotechnical and Geoenvironmental Engineering 134(8):1035–1049.
Olson, S.M., and T.D. Stark. 2002. Liquefied strength ratio from liquefaction flow failure case histories.
Canadian Geotechnical Journal 39:629–647.
Olson, S.M., and T.D. Stark. 1998. CPT based liquefaction resistance of sandy soils. Geotechnical
Earthquake Engineering and Soil Dynamics III, Geotechnical Special Publication (GSP)
75(1):325–333.
Olson, S.M., and T.D. Stark. 2002. Liquefaction strength ratio from liquefaction flow failure case
histories. Canadian Geotechnical Journal 39:629–647.
Olson, S.M., R.A. Green, and S.F. Obermeier. 2005. Engineering geologic and geotechnical analysis of
paleoseismic shaking using liquefaction effects: A major updating. Engineering Geology 76:235–
261.
Orense, R.P., T. Kiyota, S. Yamada, Y. Hosono, M. Okamura, and S. Yasuda. 2011. Comparison of
liquefaction features observed during the 2010 and 2011 Canterbury earthquakes. Seismological
Research Letters 82:905–918.
O’Rourke, T.D. 2010. Geohazards and large geographically distributed systems. Géotechnique
60(7):503–543.
O’Rourke, T.D. and S-S. Jeon. 2000. Seismic zonation for lifelines and utilities. Invited Keynote Paper on
Lifelines. In Proceedings Sixth International Conference on Seismic Zonation, Palm Springs,
California, EERI CD ROM. 35 pp.
O’Rourke, T D., S-S. Jeon, S. Toprak, M. Cubrinovski, M. Hughes, S. van Ballegooy, D. Bouziou. 2014.
Earthquake response of underground pipeline networks in Christchurch, NZ. Earthquake Spectra
30(1):183–204.
OSBGE (Oregon State Board of Geologist Examiners). 2014. Guideline for Preparing Engineering
Geologic Reports, 2nd Ed. Available at
http://www.oregon.gov/osbge/pdfs/Publications/EngineeringGeologicReports_5.2014.pdf;
accessed 29 November 2016.
Osterberg, J.O. 1952. New Piston-Type Sampler. Engineering News Record, 24 April.
O’Sullivan, C. 2011. Particulate Discrete Element Modeling: A Geomechanics Perspective. Hoboken,
NJ: Taylor & Francis.
Page, R.A., D.M. Boore, and R.F. Yerkes. n.d. The Los Angeles Dam story. Available at
http://earthquake.usgs.gov/learn/publications/la-damstory; accessed 24 September 2014.
Palmgren, A. 1924. Die lebensdauer von kugellagern [Life length of roller bearings]. Zeitschrift des
Vereins Deutscher Ingenieure 68(14):339–341.(In German).
REFERENCES 235

Pampel, F.C. 2000. Logistic Regression: A Primer. QASS No. 132. Thousand Oaks, CA: Sage
Publications.
Papathanassiou, G. 2008. LPI-based approach for calibrating the severity of liquefaction-induced failures
and for assessing the probability of liquefaction surface evidence. Engineering Geology 96:94–
104.
Papathanassiou, G., S. Pavlides, A. Ganas. 2005. The 2003 Lefkada earthquake: Field observation and
preliminary microzonation map based on liquefaction potential index for the town of Lefkada.
Engineering Geology 82:12–31.
Parra, E. 1996. Numerical modeling of liquefaction and lateral ground deformation including cyclic
mobility and dilation response in soil systems. Ph.D. diss., Rensselaer Polytechnic Institute, Troy,
NY.
Pastor, M., B. Haddad, G. Sorbino, S. Cuomo, and V. Drempetic. 2009. A depth-integrated, coupled SPH
model for flow-like landslides and related phenomena. International Journal for Numerical and
Analytical Methods in Geomechanics 33:143–172.
PEER (Pacific Earthquake Engineering Center). 2010. Guidelines for Performance Based Seismic Design
of Tall Buildings. Report No. 2010/5. Pacific Earthquake Engineering Center, College of
Engineering, University of California, Berkeley. 84 pp.
Peng, J., J. Lu, K.H. Law, and A. Elgamal. 2004. ParCYCLIC: Finite element modeling of earthquake
liquefaction response on parallel computers. International Journal for Numerical and Analytical
Methods in Geomechanics 28(12):1207–1232.
Petersen, M.D., M.P. Moschetti, P.M. Powers, C.S. Mueller, K.M. Haller, A.D. Frankel, Y. Zeng, S.
Rezaeian, S.C. Harmsen, O.S. Boyd, N. Field, R. Chen, K.S. Rukstales, N. Luco, R.L. Wheeler,
R.A. Williams, A.H. Olsen. 2014. Documentation for the 2014 update of the United States
National Seismic Hazard Maps. Available at http://purl.fdlp.gov/GPO/gpo54226; accessed 26
February 2016.
Pinheiro, J.C., and D.M. Bates. 2000. Mixed-effects models in S and S-PLUS. New York: Springer.
Pitilakis, K., K. Makropoulos, P. Bernard, F. Lemeille, H. Lyon-Caen, C. Berge-Thierry, Th. Tika, M.
Manakou, D. Diagourtas, D. Raptakis, P. Kallioglou, K. Makra, D. Pitilakis, and L.F. Bonilla.
2004. The Corinth Gulf Soft Soil Array (CORSSA) to study site effects. Comptes Rendus
Geoscience 336:353–365.
Pohari, M., M. Blas, and S. Turk. 2004. Comparison of logistic regression and linear discriminant
analysis: A simulation study. Metodološki zvezki 1(1):143–161.
Polito, C.P., and J. Martin. 2001. Effects of nonplastic fines on the liquefaction resistance of sands.
Journal of Geotechnical and Geoenvironmental Engineering 127(5):408–415.
Polito, C.P., R.A. Green, and J. Lee. 2008. Pore pressure generation models for sands and silty soils
subjected to cyclic loading. Journal of Geotechnical and Geoenvironmental Engineering
134(10):1490–1500.
Ponthot, J.-P., and T. Belytschko. 1998. Arbitrary Lagrangian-Eulerian formulation for element-free
Galerkin method. Computer Methods in Applied Mechanics and Engineering 152:19–46.
POLA (Port of Los Angeles). 2010. Code for Seismic Design, Upgrade, and Repair of Container
Wharves. San Pedro, CA: City of Los Angeles, Harbor Department.
POLB (Port of Long Beach). 2012. Wharf Design Criteria. Long Beach, CA: Port of Long Beach. Pp.
69–84.
Potapov, A.V., M.L. Hunt, and C.S. Campbell. 2001. Liquid-solid flows using smoothed particle
hydrodynamics and the discrete element method. Powder Technology 116:204–213.
Poulos, S.J. 1981. The steady state of deformation. Journal of Geotechnical Engineering Division
107(GT5):553–562.
Prevost, J.H. 1980. Mechanics of continuous porous media. International Journal of Engineering Science
18(6):787–800.
Prevost, J.H. 1985a. A simple plasticity theory for frictional cohesionless soils. Soil Dynamics and
Earthquake Engineering 4(1):9–17.
236 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Prevost, J.H. 1985b. Wave propagation in fluid-saturated porous media: An efficient finite element
procedure. Soil Dynamics and Earthquake Engineering 4(4):183–202.
Prevost, J.H., A.M. Abdel-Ghaffar, and S.J. Lacy. 1985. Nonlinear dynamic analyses of an earth dam.
Journal of Geotechnical Engineering 111(7): 882–897.
Pyke, R.M. 2000. TESS: A Computer Program for Nonlinear Ground Response Analyses. Lafayette, CA:
TAGA Systems & Software.
Pyke, R.M., C.K. Chan, and H.B. Seed. 1974. Settlement and liquefaction of sands under multi-
directional shaking. Report Number UCB/EERC-74/02. Earthquake Engineering Research
Center, University of California, Berkeley. 46 pp.
Qadimi, A., and M. Coop. 2007. The undrained cyclic behaviour of a carbonate sand. Géotechnique
557(9):739–750.
Quigley, M.C., S. Bastin, and B.A. Bradley. 2013. Recurrent liquefaction in Christchurch, New Zealand,
during the Canterbury earthquake sequence. Geology 41(4):419–422.
Randles, P.W., and L.D. Libersky. 1996. Smoothed particle hydrodynamics: Some recent improvements
and applications. Computer Methods in Applied Mechanics and Engineering 139(1):375–408.
Raschke, S.A. and R.D. Hryciw. 1997. Vision cone penetrometer (VisCPT) for direct subsurface soil
observation. Journal of Geotechnical and Geoenvironmental Engineering 123(11):1074–1076.
Rathje, E.M., and G. Antonakos. 2011. A unified model for predicting earthquake-induced sliding
displacements of rigid and flexible slopes. Engineering Geology 122:51–60,
doi:10.1016/j.enggeo.2010.12.004.
Rathje, E.M., and K. Franke. 2015. Remote sensing for geotechnical earthquake reconnaissance. In
Proceedings of the 6th International Conference on Earthquake Geotechnical Engineering,
Christchurch, New Zealand.
Rathje, E. M., and M.C. Ozbey. 2006. Site-specific validation of random vibration theory-based seismic
site response analysis. Journal of Geotechnical and Geoenvironmental Engineering 132(7):911 –
922.
Rathje, E.M., W.-J. Chang, and K.H. Stokoe II. 2005. Development of an in situ dynamic liquefaction.
Geotechnical Testing Journal 28(1):50–60.
Rathje, E.M., W.-J. Chang, K.H. Stokoe II, and B.R. Cox. 2004. Evaluation of ground strain from in situ
dynamic testing. Paper No. 3099 in 13th World Conference on Earthquake Engineering, August,
Vancouver, Canada.
Rauch F.A., and R.J. Martin II. 2000. EPOLLS model for predicting average displacements on lateral
spreads. Journal of Geotechnical and Geoenvironmental Engineering 126(4):360–371.
Reese, L.C., W.R. Cox, and F.D. Koop. 1974. Analysis of Laterally Loaded Piles in Sand. Pp. 95-104 in
Proceedings of the Offshore Technology Conference, 6–8 May, Houston, Texas. Paper No. OTC
2080.
Reimer, M. 2007. Cyclic Simple Shear Testing to Assess Likely Settlements for BART Extension. Final
report prepared for Kleinfelder, Inc. for the Silicon Valley Berryessa Extension Project, Bay Area
Rapid Transit System. 41 pp.
Riemer, M.F., W.B. Gookin, J.D. Bray, and I. Arango. 1994. Effects of Loading Frequency and Control
on the Liquefaction Behavior of Clean Sands. Geotechnical Engineering Report No. UCB/GT/94-
07. Department of Civil and Environmental Engineering, University of California, Berkeley.
Robertson, P.K. 1990. Soil classification using CPT. Canadian Geotechnical Journal 27(1):151–158.
Robertson, P.K. 2009. Performance based earthquake design using the CPT. Performance Based Design.
Pp. 3–20 in Earthquake Geotechnical Engineering: From Case History to Practice, edited by T.
Kokusho, Y. Tsukamoto, and M. Yoshimine. London: Taylor & Francis.
Robertson, P.K. 2010. Evaluation of flow liquefaction and liquefaction strength using the cone
penetration test. Journal of Geotechnical and Geoenvironmental Engineering 136(6):842–853.
Robertson, P.K., and K.L. Cabal. 2014. Guide to Cone Penetration Testing for Geotechnical Engineering,
6th Ed. Signal Hill, CA: Gregg Drilling & Testing, Inc. Pp. 4–5.
REFERENCES 237

Robertson, P.K., and R.G. Campanella. 1985. Liquefaction Potential of Sands Using the CPT. Journal of
Geotechnical Engineering 111:384–403.
Robertson, P.K., and C.E. Wride. 1998. Evaluating cyclic liquefaction potential using the cone
penetration test. Canadian Geotechnical Journal 35(3):442–459.
Robertson, P.K., C.E. Wride (Fear), B.R. List, U. Atukorala, K.W. Biggar, P.M. Byrne, R.G. Campanella,
D.C. Cathro, D.H. Chan, K. Czajewski, W.D.L. Finn, W.H. Gu, Y. Hammamji, B.A. Hofmann,
J.A. Howie, J. Hughes, A.S. Imrie, J.-M. Konrad, A. Küpper, T. Law, E.R.F. Lord, P.A..
Monahan, N.R. Morgenstern, R. Phillips, R. Piché, H.D. Plewes, D. Scott, D.C. Sego, J.
Sobkowicz, R.A. Stewart, B.D. Watts, D.J. Woeller, T.L. Youd, and Z. Zavodni. 2000. The
CANLEX project: Summary and conclusions. Canadian Geotechnical Journal 37:563–591.
Rogers, N., K. Williams, M. Jacka, S. Wallace, and J. Leeves. 2014. Geotechnical aspects of disaster
recovery planning in residential Christchurch and surrounding districts affected by liquefaction.
Earthquake Spectra 30:493–512, doi:10.1193/021513EQS029M.
Rollins, K.M., T.M. Gerber, J.D. Lane, and S.A. Ashford. 2005. Lateral resistance of a full-scale pile
group in liquefied sand. Journal of Geotechnical and Geoenvironmental Engineering
131(1):115–125.
Salgado, R., J. Mitchell, and M. Jamiolkowski. 1997. Cavity expansion and penetration resistance in and.
Journal of Geotechnical and Geoenvironmental Engineering 123(4):344–354.
Saygili, G., and E.M. Rathje. 2008. Empirical predictive models for earthquake-induced sliding
displacements of slopes. Journal of Geotechnical and Geoenvironmental Engineering
134(6):790–803.
SCDOT (South Carolina Department of Transportation). 2010. SCDOT Geotechnical Design Manual.
Chapter 13, Geotechnical Seismic Hazards. Available at
http://www.scdot.org/doing/technicalpdfs/geotechnicaldesign/chapter%2013%20geotechnical%2
0seismic%20hazards%20-%2005052010.pdf; accessed November 2,92016.
Schmertmann, J.H. 1991. The mechanical aging of soils. Journal on Geotechnical Engineering
117(9):1288–1330.
Schofield, A., and P. Wroth. 1968. Critical State Soil Mechanics. New York: McGraw-Hill.
Seed, H.B. 1968. Landslides during earthquakes due to soil liquefaction. Journal of the Soil Mechanics
and Foundations Division 94:1053–1122.
Seed, H.B. 1979. Soil liquefaction and cyclic mobility evaluation for level ground during earthquakes.
Journal of the Geotechnical Engineering Division 105(GT2):201–255.
Seed H.B. 1983. Earthquake-resistant design of earth dams. Pp. 41–64 in Proceedings of Symposium on
Seismic Design of Embankments and Caverns, ASCE, Philadelphia, Pennsylvania.
Seed, H.B. 1987. Design problems in soil liquefaction. Journal of Geotechnical Engineering 113(8):827–
845.
Seed, H.B., and L.F. Harder Jr. 1990. SPT-based analysis of cyclic pore pressure generation and
undrained residual strength. Pp. 351–376 in H. B. Seed Memorial Symposium, Vol. 2, edited by
J.M. Duncan. Vancouver, Canada: Bi-Tech Publishers Ltd.
Seed, H.B., and I.M. Idriss. 1970. Soil moduli and damping factors dynamic response analyses. Report
EERC 70-10. University of California, Berkeley.
Seed, H.B., and I.M. Idriss. 1971. Simplified procedure for evaluating soil liquefaction potential. Journal
of Geotechnical Engineering Division 97(9):1249–1273.
Seed, H.B., and I.M. Idriss. 1982. Ground motions and soil liquefaction during earthquakes. Oakland,
CA: Earthquake Engineering Research Institute.
Seed, H.B., K.L. Lee, I.M. Idriss, and F. Makdisi. 1973. Analysis of the Slides in the San Fernando Dams
During the Earthquake of Feb. 9, 1971. Earthquake Engineering Research Center, University of
California, Berkeley.
Seed, H.B., K.L. Lee, I.M. Idriss, and F.I. Makdisi. 1975a. The slides in the San Fernando Dams during
the earthquake of February 9, 1971. Journal of the Geotechnical Engineering Division 101:651–
688.
238 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Seed, H.B., I.M. Idriss, F. Makdisi, and N. Banerjee. 1975b. Representation of Irregular Stress Time
Histories by Equivalent Uniform Stress Series in Liquefaction Analyses. Report No. EERC 75-29.
Earthquake Engineering Research Center, University of California, Berkeley.
Seed, H.B., P. Martin, and J. Lysmer. 1976. Pore water pressure changes during soil liquefaction. Journal
of the Geotechnical Engineering 102(4):323–345.
Seed, H.B., K. Tokimatsu, L.F. Harder Jr., and R. Chung. 1984. Influence of SPT Procedures in Soil
Liquefaction Resistance Evaluations. Report No. UCB/EERC-84/15. Earthquake Engineering
Research Center, University of California, Berkeley.
Seed, H.B., K. Tokimatsu, L.F. Harder Jr., and R. Chung. 1985. Influence of SPT procedures in soil
liquefaction resistance evaluations. Journal of Geotechnical Engineering 111(12):1425–1445.
Seed, H., R. Wong, I. Idriss, and K. Tokimatsu. 1986. Moduli and damping factors for dynamic analyses
of cohesionless soils. Journal of Geotechnical Engineering 112(11):1016–1032.
Seed, H.B., R.B. Seed, L.F. Harder, and H.L. Jong. 1988a. Re-evaluation of the Slide in the Lower San
Fernando Dam in the 1971 San Fernando Earthquake. Report No. UCB/EERC-88/04.
Earthquake Engineering Research Center, University of California, Berkeley.
Seed, H.B., R.B. Seed, F. Schlosser, F. Blondeau, and I. Juran. 1988b. The Landslide at the Port of Nice
on October 16, 1979. Report No. UCB/EERC-88/10. Earthquake Engineering Research Center,
University of California, Berkeley.
Seed, R.B., K.O. Cetin, R.E.S. Moss, A.M. Kammerer, J. Wu, J.M. Pestana, M.F. Riemer, R.B. Sancio,
J.D. Bray, R.E. Kayen, and A. Faris. 2003. Recent Advances in Soil Engineering: A Unified and
Consistent Framework. Earthquake Engineering Research Center Report No. EERC 2003-06.
University of California, Berkeley.
Semple, R. 2013. Problems with liquefaction criteria and their application in Australia. Australian
Geomechanics 48(3):15–34.
Shahir, H., and A. Pak. 2010. Estimating liquefaction induced settlement of shallow foundations by
numerical approach. Computers and Geotechnics 37(3):267–279.
Shamoto, Y., J.-M Zhang, and K. Tokimatsu. 1998. Methods for evaluating residual post-liquefaction
ground settlement and horizontal displacement. Soils and Foundations 38(Special Issue No.
2)69–83.
Shuttle, D.A., and J. Cunning. 2007. Liquefaction potential of silts from CPTu. Canadian Geotechnical
Journal 44(1):1–19.
Siegel, T.C. 2013. Liquefaction mitigation synthesis report. DFI (Deep Foundations Institute) Journal
7:13–31.
Sieh, K.E. 1978. Prehistoric large earthquakes produced by slip on the San Andreas Fault at Pallett Creek,
California. Journal of Geophysical Research 83:3907–3939, doi:10.1029/JB083iB08p03907.
Silva, V., H. Crowley, M. Pagani, D. Monelli, and R. Pinho. 2014. Development of the OpenQuake
engine, the Global Earthquake Model’s open-source software for seismic risk assessment.
Natural Hazards 72(3):1409–1427.
Silver, M.L., and H.B. Seed. 1971. Volume changes in sands during cyclic loading. Journal of the Soil
Mechanics and Foundation Division 97(SM9):1171–1182.
Simo, J.C., and T.J.R. Hughes. 1998. Computational Inelasticity. New York: Prentice Hall.
Sims, J.D., and C.D. Garvin. 1995. Recurrent liquefaction induced by the 1989 Loma Prieta earthquake
and 1990 and the 1991 aftershocks: Implications for paleoseismicity studies. Bulletin of the
Seismological Society of America 85(1):51–65.
Sitar, N. ed. 1995. Geotechnical reconnaissance of the effects of the January 17, 1995. Hyogoken-Nanbu
earthquake, Japan. Report No. UCB/EERC-95/01. Earthquake Engineering Research Center,
University of California, Berkeley.
Skempton, A.W. 1960. Effective stresses in soil, concrete, and rocks. Pp. 4–16 in Proceedings of the
Conference on Pore Pressure and Suction in Soils. London: Butterworths.
REFERENCES 239

Skempton, A.W. 1986. Standard penetration test procedures and the effects in sands of overburden
pressure, relative density, particle size, aging, and overconsolidation. Géotechnique 36(3):425–
447.
Smyrou, E., P. Tasiopoulou, I. Engin Bal, and G. Gazetas. 2011. Ground motions versus geotechnical and
structural damage in the February 2011 Christchurch earthquake. Seismological Research Letters
82:882–892.
Sonmez, H. 2003. Modification of the liquefaction potential index and liquefaction susceptibility mapping
for a liquefaction-prone area (Inegol, Turkey). Environmental Geology 44:862–871.
Stafford, P.J., and Bommer, J.J. 2009. Empirical equations for prediction of the equivalent number of
cycles of earthquake ground motion. Soil Dynamics and Earthquake Engineering 29:1425–1436.
Starbird, K., and L. Palen. 2013. Working and Sustaining the Virtual Disaster Desk. Pp. 491–502 in
Proceedings of the ACM Conference on Computer Supported Cooperative Work, San Antonio,
Texas.
Stark, T.D., and G. Mesri. 1992. Undrained shear strength of liquefied sands for stability analysis.
Journal of Geotechnical Engineering 118(11):1727–1747.
Stewart, J.P., and A.O.L. Kwok. 2008. Nonlinear seismic ground response analysis: Code usage protocols
and verification against vertical array data. In Proceedings of the ASCE conference of
Geotechnical Earthquake Engineering and Soil Dynamics IV, , Sacramento, California.
Geotechnical Special Publication 181. Reston, VA: American Society of Civil Engineers.
Stark, T.D., and S.M. Olson. 1995. Liquefaction resistance using CPT and field case histories. Journal of
Geotechnical Engineering 121(12):856–869.
Stark, T.D., S.M. Olson, S.L. Kramer, and T.L. Youd. 1999. Shear strength of liquefied soil. ASCE
Geotechnical Special Publication 1(75):313–324.
Stewart, J.P., S.L. Kramer, D.Y. Kwak, M.W. Greenfield, R.E. Kayen, K. Tokimatsu, J.D. Bray, C.Z.
Beyzaei, M. Cubrinovski, T. Sekiguchi, S. Nakai, and Y. Bozorgnia. 2015. PEER-NGL project:
Open source global database and model development for the next-generation of liquefaction
assessment. In Proceedings of the 6th International Conference on Earthquake Geotechnical
Engineering, Christchurch, New Zealand. 23 pp.
Stokoe, K.H., and J.C. Santamarina. 2000. Seismic-wave-based testing in geotechnical engineering.
International Society for Rock Mechanics. Document No. ISRM-IS-2000-38. In Proceedings of
the ISRM International Symposium, 19–24 November, Melbourne, Australia. 47 pp.
Stokoe, K.H., II, G.J. Rix, I. Sanchez-Salinero, R.D. Andrus, and Y.J. Nok. 1988. Liquefaction of
gravelly soils during the 1983 Borah Peak, ID earthquake. Pp. 183–188 in Proceedings of the 9th
World Conference in Earthquake Engineering, Tokyo, Japan.
Stokoe, K.H., II, J.N. Roberts, S. Hwang, B.R. Cox, F.Y. Menq, and S. van Ballegooy. 2014a.
Effectiveness of inhibiting liquefaction triggering by shallow ground improvement methods:
Initial field shaking trials with T-Rex at one site in Christchurch, New Zealand. In Soil
Liquefaction during Recent Large-Scale Earthquakes, edited by R.P. Orense, I. Towhata and N.
Chouw. London: Taylor & Francis.
Stokoe, K.H., II, J.N. Roberts, B.R. Cox, S. Hwang, F.Y. Menq, and S. van Ballegooy. 2014b.
Effectiveness of shallow ground improvements to inhibit liquefaction triggering: Methodology
and analysis of field trials using controlled-source, dynamic loading with T-Rex. Submitted to
Soil Dynamic and Earthquake Engineering.
Sukop, M.C. and D.T. Thorne. 2006. Lattice Boltzmann Modeling: An Introduction for Geoscientists and
Engineers. New York: Springer.
Sulsky, D., Z. Chen, and H.L. Schreyer. 1994. A particle method for history-dependent materials.
Computer Methods in Applied Mechanics and Engineering 118:179–196.
Sun, W.C., J.E. Andrade, J.W. Rudnicki, and P. Eichhubl. 2011a. Connecting microstructural attributes
and permeability from 3D tomographic images of in situ shear-enhanced compaction bands using
multiscale computations. Geophysical Research Letters 38:L10302.
240 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Sun, W.C., J.E. Andrade, and J.W. Rudnicki. 2011b. Multiscale method for characterization of porous
microstructures and their impact on macroscopic effective permeability. International Journal for
Numerical Methods in Engineering 88(12):1260–1279.
Sutterer, K.G., D.J. Frost, and J.-L.A. Chameau. 1996. Polymer impregnation to assist undisturbed
sampling of cohesionless soils. Journal of Geotechnical Engineering 122(3):209–215.
Suzuki, Y., K. Tokimatsu, K. Koyamada, Y. Taya, and Y. Kubota. 1995. Field correlation of soil
liquefaction based on CPT data. Pp. 583–588 in Proceedings of the International Symposium on
Cone Penetration Testing (CPT ’95), 2 October, Linkoping, Sweden.
Sy, A., and R.G. Campanella. 1994. Becker and standard penetration tests (BPT-SPT) correlations with
consideration of casing friction. Canadian Geotechnical Journal 31:343–356.
Takada, K., and B.F. Atwater. 2004. Evidence for liquefaction identified in peeled slices of Holocene
deposits along the lower Columbia River, Washington. Bulletin of the Seismological Society of
America 94:550–575.
Talwani, P., and W. Schaeffer. 2001. Recurrence rates of large earthquakes in the South Carolina coastal
plain based on paleoliquefaction data. Journal of Geophysical Research-Solid Earth 106:6621–
6642, doi:10.1029/2000JB900398.
Tatsuoka, F., T. Sasaki, and S. Yamada. 1984. Settlement in saturated sand induced by cyclic undrained
simple shear. Pp. 95–102 in Proceeding of the 8th World Conference on Earthquake Engineering,
San Francisco, California.
Taylor, M. L., M. Cubrinovski, and B.A. Bradley. 2012. Characterisation of ground conditions in the
Christchurch central business district. Australian Geomechanics. 47(4):43-58.
Terzaghi, C. 1925. Principles of soil mechanics. Engineering News-Record 95:19–27.
Tinti, S., A. Armigliato, A. Manucci, G. Pagnoni, F. Zaniboni, A. Yalciner, and Y. Altinok. 2006. The
generating mechanisms of the August 17, 1999 Izmit Bay (Turkey) tsunami: Regional (tectonic)
and local (mass instabilities) causes. Marine Geology 225:311–330.
Tobita, T., G.-C. Kang, and S. Iai. 2012. Estimation of liquefaction-induced manhole uplift displacements
and trench-backfill settlements. Journal of Geotechnical and Geoenvironmental Engineering
138(4):491–499 .
Tohno, I., and Y. Shamoto. 1986. Liquefaction damage to the ground during the 1983 Nihonkai-Chubu
earthquake in Aomori Prefecture. Natural Disaster Science 8(1):85–116.
Tokimatsu, K., and Y. Asaka. 1998. Effects of liquefaction induced ground displacements on pile
performance in the 1995 Hyogoken-Nambu earthquake. Soils and Foundations (Special
Issue):163–177.
Tokimatsu, K., and H.B. Seed. 1987. Evaluation of settlements in sands due to earthquake shaking.
Journal of the Geotechnical Engineering Division 113(8):861–878.
Tokimatsu, K. and Y. Yoshimi. 1983. Empirical correlation of soil liquefaction based on SPT N-value
and fines content. Soils and Foundations 23(4):56–74.
Tokimatsu, K., Yamazako, T., and Yoshimi, Y. 1986. Soil liquefaction evaluations by elastic shear
moduli. Soils and Foundations 26(1):25–35.
Tonkin & Taylor Ltd. 2013. Liquefaction Vulnerability Study. Report to the Earthquake Commission,
prepared by S. van Ballegooy and P. Malan. T&T Ref: 52020.0200/v1.0. Available at
http://www.eqc.govt.nz/sites/public_files/documents/liquefaction-vulnerability-study-final.pdf;
accessed 29 November 2016.
Toprak, S., and T. Holzer. 2003. Liquefaction potential index: Field assessment. Journal of Geotechnical
and Geoenvironmental Engineering 129(4):315–322.
Towhata, I., and K. Ishihara. 1985a. Modeling soil behavior under principal axes rotation. Pp. 523–530 in
Proceedings of Fifth International Conference on Numerical Methods in Geomechanics,.
Nagoya, Japan.
Towhata, I., and K. Ishihara. 1985b. Shear work and pore water pressure in undrained shear. Soils and
Foundations 25(3):73-84.
REFERENCES 241

Travasarou, T., J.D. Bray, and N.A. Abrahamson. 2003. Empirical attenuation relationship for Arias
Intensity. Earthquake Engineering and Structural Dynamics 32:1133–1155.
Troncoso, J., K. Ishihara, and R. Verdugo. 1988. Aging effects on cyclic shear strength of tailing
materials. Pp. 121–126 in Proceedings of the 9th World Conference on Earthquake
Engineering¸Vol. 3, Tokyo, Japan. .
Tsuji, Y., T. Kawaguchi, and T. Tanaka. 1993. Discrete particle simulation of two-dimensional fluidized
bed. Powder Technology 77(1):79–87.
Tuttle, M.P., and R.D. Hartleb. 2012. CEUS paleoliquefaction database, uncertainties associated with
paleoliquefaction data, and guidance for seismic source characterization. Pp. E-i–E-135 in
Technical Report: Central and Eastern United States Seismic Source Characterization for
Nuclear Facilities, Vol. 2. NUREG-2115. Washington, DC: Electric Power Research Institute,
U.S. Department of Energy, and U.S. Nuclear Regulatory Commission. Available at
http://www.ceus-ssc.com/PDF/AppendixE.pdf; accessed 29 November 2016.
Tuttle, M.P., and L. Seeber. 1991. Historic and prehistoric earthquake-induced liquefaction in Newbury,
Massachusetts. Geology 19:594–597, doi:10.1130/0091-
7613(1991)019<0594:HAPEIL>2.3.CO;2.
Unutmaz, B., and Cetin K. O. 2012. Post cyclic settlement and tilting potential of mat foundations. Soil
Dynamics and Earthquake Engineering 43:271–286.
Urzúa, A., and J.T. Christian. 2013. Sliding displacements due to subduction-zone earthquakes.
Engineering Geology 166:237–244.
USBR (U.S. Bureau of Reclamation). 2011. Design Standard No. 13: Embankment Dams. Available at
http://www.usbr.gov/tsc/techreferences/designstandards-
datacollectionguides/designstandards.html; accessed 29 November 2016.
USNRC (U.S. Nuclear Regulatory Commission). 1975. Reactor Safety Analysis: An Assessment of
Accident Risks in U.S. Commercial Nuclear reactors. WASH-1400 (NUREG-75/104).
Washington, DC: U.S. Nuclear Regulatory Commission.
USNRC. 2003. Regulatory Guide 1.198 Procedures and Criteria for Assessing Seismic Soil Liquefaction
at Nuclear Power Plant Sites. Washington, DC: U.S. Nuclear Regulatory Commission, Office of
Nuclear Regulatory Research. Available at
http://permanent.access.gpo.gov/lps106679/doccontent.pdf; accessed 29 November 2016.
Vaid, Y.P., and S. Sivathayalan. 1996. Static and cyclic liquefaction potential of Fraser Delta sand in
simple shear and triaxial tests. Canadian Geotechnical Journal 33(2):281–289.
van Ballegooy, S., P.J. Malan, M.E. Jacka, V.I.M.F. Lacrosse, J.R. Leeves, and J.E. Lyth. 2012. Methods
for Characterizing Effects of Liquefaction in Terms of Damage Severity. In Proceedings of the
15th World Conference on Earthquake Engineering, Lisbon, Portugal.
van Ballegooy, S., P. Malan, V. Lacrosse, M.E. Jacka, M. Cubrinovski, J.D. Bray, T.D. O’Rourke, S.A.
Crawford, and H. Cowan. 2014a. Assessment of liquefaction-induced land damage for residential
Christchurch. Earthquake Spectra 30(1):31–55, doi:10.1193/031813EQS070M.
van Ballegooy, S., K. Berryman, B. Deam, and M. Jacka. 2014b. Repeated Major Episodes of Tectonic
Deformation, Lateral Spread and Liquefaction in Christchurch During the Canterbury
Earthquake Sequence of 2010-2011. IAEG XII Congress, Torino, Italy.
van Ballegooy, S., R.A. Green, J. Lees, F. Wentz, and B.W. Maurer. 2015. Assessment of various CPT
based liquefaction severity index frameworks relative to the Ishihara (1985) H1-H2 boundary
curves. Soil Dynamics and Earthquake Engineering 79(Part B):347–364.
Verdugo, R. 2009. Amplification phenomena observed in downhole array records generated on a
subductive environment. Physics of the Earth and Planetary Interiors 175:63–77.
Vucetic, M. 1986. Pore pressure buildup and liquefaction at level sand sites during earthquakes. Ph.D.
diss., Rensselaer Polytechnic Institute, Troy, NY.
Wang, C.H. 2003. Prediction of the residual strength of liquefied soils. Ph.D. diss., University of
Washington, Seattle. 457 pp.
242 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Wang, J.-H., Moran, K., and Baxter, C.D.P. 2006. Correlation between cyclic resistance ratios of intact
and reconstituted offshore saturated sands and silts with the same shear wave velocity. Journal of
Geotechnical and Geoenvironmental Engineering, ASCE 132(12):1574–1580.
Wang, J.N, and E. Kavazanjian, Jr. 1989. Pore pressure development during non-uniform cyclic loading.
Soils and Foundations 29(2):1–14.
Wang, S.T., and L.C. Reese. 1998. Design of pile foundations in liquefied soils. Pp. 1331–1343 in
Geotechnical Earthquake Engineering and Soil Dynamics III, Vol. 2, Geotechnical Special
Publication No. 75, edited by P. Dakulos and M. Yegian. Reston, VA: American Society of Civil
Engineers.
Wang, W.S. 1979. Some Findings in Soil Liquefaction. Beijing: Water Conservancy and Hydroelectric
Power Scientific Research Institute.
Weber, J.P. 2015. Engineering evaluation of post-liquefaction strength. Ph.D. diss., University of
California, Berkeley. 158 pp. Available at
http://digitalassets.lib.berkeley.edu/etd/ucb/text/Weber_berkeley_0028E_15600.pdf; accessed 26
October 2016.
Wellmann, C., C. Lillie, and P. Wriggers. 2008. Homogenization of granular material modeled by a three-
dimensional discrete element method. Computers and Geotechnics 35(3):394–405.
White, J.A., R.I. Borja, and J.T. Fredrich. 2006. Calculating the effective permeability of sandstone with
multiscale lattice Boltzmann/finite element simulations. Acta Geotechnica 1(4):195–209.
Whitman, R.V. 1971. Resistance of soil to liquefaction and settlement. Soils and Foundations 11(4):59–
68.
Whitman, R.V., J.W. Reed, and S.T. Hong. 1974. Earthquake damage probability matrices. Paper no
2531 in Proceedings of the Fifth World Conference on Earthquake Engineering, International
Association for Earthquake Engineering, Rome, Italy.
Wilson, D.W. 1998. Soil-pile-superstructure interaction at soft and liquefying sites. Ph.D. Dissertation.
University of California, Davis.
Wilson, D.W., R.W. Boulanger, and B.L. Kutter. 2000. Observed seismic lateral resistance of liquefying
sand. Journal of Geotechnical and Geoenvironmental Engineering 126 (10):898–906.
Wood, C. M., B. R. Cox, L. M. Wotherspoon, and R. A. Green. 2011. Dynamic site characterization of
Christchurch strong motion stations. Bulletin of the NZSEE 44(4):195–204.
Wotherspoon, L.M., M.J. Pender, and R.P. Orense. 2012. Relationship between observed liquefaction at
Kaiapoi following the 2010 Darfield earthquake and former channels of the Waimakariri River.
Engineering Geology 125:45–55.
Wotherspoon, L.M., R.P. Orense, M. Jacka, R.A. Green, B.R. Cox, and C.M. Wood. 2014. Seismic
performance of improved ground sites during the 2010-2011 Canterbury earthquake sequence.
Earthquake Spectra 30(1):111–119.
Wu, J. 2002. Liquefaction triggering and post-liquefaction deformation of Monterey 0/30 sand under uni-
directional cyclic simple shear loading. Ph.D. diss., University of California, Berkeley.
Wu, J., R.B. Seed, and J.M. Pestana. 2003. Liquefaction triggering and post liquefaction deformations of
Monterey 0/30 sand under uni-directional cyclic simple shear loading. Geotechnical Engineering
Research Report No. UCB/GE-2003/01. University of California, Berkeley.
Xu, J., J. Bielak, O. Ghattas, and J. Wang. 2003. Three dimensional nonlinear seismic ground motion
modeling in inelastic basins. Physics of the Earth and Planetary Interiors 137:81–95.
Yalcin, A., C. Gokceoglu, and H. Sonmez. 2008. Liquefaction severity map for Aksaray city center.
Natural Hazards and Earth Science Systems 8:641–649.
Yanagisawa, E., and T. Sugano. 1994. Undrained shear behaviors of sand in view of shear work. Pp. 155–
158 in Performance of Ground and Soil Structures during Earthquakes: Thirteenth International
Conference on Soil Mechanics and Foundation Engineering, 5–10 January, New Delhi. Boca
Raton, FL: CRC Press.
Yang, Z., and A. Elgamal. 2002. Influence of permeability on liquefaction induced shear deformation.
Journal of Engineering Mechanics 128(7):720–729.
REFERENCES 243

Yang, Z., A. Elgamal, and E. Parra. 2003. Computational model for cyclic mobility and associated shear
deformation. Journal of Geotechnical and Geoenvironmental Engineering 129(12):1119–1127.
Yang, Z., J.L. Lu, and A. Elgamal. 2008. OpenSees Soil Models and Solid-Fluid Fully Coupled Elements.
User’s Manual: 2008 ver 1.0. Department of Structural Engineering, University of California,
San Diego.
Yasuda, S., and I. Tohno. 1988. Sites of reliquefaction caused by the 1983 Nihonkai-Chubu earthquake.
Soils and Foundations 28(2):61–72.
Yegian, M.K., and B. Vitelli. 1981. Analysis for liquefaction: Empirical approach. Pp. 173–177 in the
Proceedings of the International Conference on Recent Advances in Geotechnical Earthquake
Engineering and Soil Dynamics, 26 April–3 May, St. Louis, Missouri. Available at
http://nuweb5.neu.edu/myegian/wp-content/uploads/2015/01/Liquefaction-Analysis-Empirical-
Approach.pdf; accessed 26 October 2016.
Yegian, M.K., E. Eseller-Bayat, A. Alshawabkeh, and S. Ali. 2007. Induced-partial saturation for
liquefaction mitigation: Experimental investigation. Journal of Geotechnical and
Geoenvironmental Engineering 133(4):372–380.
Yoshimi, Y., M. Hatanaka, and H. Oh-Oka. 1978. Undisturbed sampling of saturated sands by freezing.
Soils and Foundation 8(3):59–73.
Yoshimine, M., and K. Ishihara. 1998. Flow potential of sand during liquefaction. Soils and Foundations
38(3):189–198.
Yoshimoto, N., R. Orense, M. Hyodo, and Y. Nakata. 2014. Dynamic behavior of granulated coal ash
during earthquakes. Journal of Geotechnical and Geoenvironmental Engineering
140(2):04013002-1–11, doi:10.1061/(ASCE)GT.1943-5606.0000986.
Youd, T.L. 1971. Landsliding in the vicinity of the Van Norman Lakes. Pp. 105–109 in The San
Fernando, California, Earthquake of February 9, 1971: A preliminary report published jointly by
the U.S. Geological Survey and the National Oceanic and Atmospheric Administration. U.S.
Geological Survey Professional Paper 733. Washington, DC: U.S. Department of the Interior,
U.S. Department of Commerce, National Oceanic and Atmospheric Administration. Available at
http://pubs.usgs.gov/pp/0733/report.pdf; accessed November 29, 2016.
Youd, T.L. 1972. Compaction of sands by repeated shear straining. Journal of the Soil Mechanics and
Foundation Division 98(SM7):709–725.
Youd, T.L. 1973. Ground movements in Van Norman Lake vicinity during San Fernando earthquake. Pp.
197–203 in San Fernando, California Earthquake of February 9, 1971: Volume III, Geological
and Geophysical Studies. Washington, DC: U.S. Department of Commerce, National
Oceanographic and Atmospheric Administration.
Youd, T.L. 1984. Geologic Effects–Liquefaction and Associated Ground Failure. In Proceedings of the
Geologic and Hydrologic Hazards Training Program. USGS Open File Report 84-760. Reston,
VA: U.S. Geological Survey.
Youd, T.L. 1991. Mapping of earthquake-induced liquefaction for seismic zonation. Pp. 111–147 in
Proceedings of the Fourth International Conference on Seismic Zonation, Vol. 1. Stanford, CA:
Earthquake Engineering Research Institute.
Youd, T.L. 2003. Liquefaction mechanisms and induced ground failure. Pp. 1159–1174 in International
Handbook on Earthquake and Engineering Seismology, edited by W.H.K. Lee, H. Kanamori,
P.C. Jennings, and C. Kisslinger. Amsterdam, the Netherlands: Academic Press.
Youd, T.L., and B.L. Carter. 2005. Influence of soil softening and liquefaction on spectral acceleration.
Geotechnical and Geoenvironmental Engineering 131:811–825.
Youd, T.L., and C.T. Garris, 1995. Liquefaction-induced ground-surfacedisruption. Journal of
Geotechnical and Geoenvironmental Engineering 121(11):805–809.
Youd, T.L., and S.N. Hoose. 1977. Liquefaction susceptibility and geologic setting. Pp. 2189–2194 in
Proceedings of the Sixth World Conference on Earthquake Engineering, Vol. 3. Englewood
Cliffs, NJ: Prentice Hall.
244 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Youd, T.L., and S.N. Hoose. 1978. Historic Ground Failures in Northern California Triggered by
Earthquakes. U.S. Geological Survey Professional Paper 993. 177 pp. Available at
http://pubs.usgs.gov/pp/1978/pp0993/pp993_text.pdf; accessed 29 November 2016.
Youd, T.L., and I.M. Idriss. 1997. Proceedings of the NCEER Workshop on Evaluation of Liquefaction
Resistance of Soils. Buffalo, NY: National Center for Earthquake Engineering Research.
Youd, T.L., and D.K. Keefer. 1994. Liquefaction during the 1977 San Juan Province, Argentina
earthquake (Ms = 7.4). Engineering Geology 37:211–233.
Youd, T.L., and S.K. Noble. 1997a. Liquefaction criteria based on statistical and probabilistic analyses.
Pp. 201–205 in Proceedings of the NCEER Workshop on Evaluation of Liquefaction Resistance
of Soils, 5–6 January, Salt Lake City, Utah. Buffalo, NY: National Center for Earthquake
Engineering Research.
Youd, T.L., and S.K. Noble. 1997b. Magnitude scaling factors. Pp. 149–165 in Proceedings of the
NCEER Workshop on Evaluation of Liquefaction Resistance of Soils, 5–6 January, Salt Lake
City, Utah. Buffalo, NY: National Center for Earthquake Engineering Research.
Youd, T.L., and D.M. Perkins. 1978. Mapping liquefaction-induced ground failure potential. Journal of
the Geotechnical Engineering Division, ASCE 104(GT4):433–446.
Youd, T.L., E.L. Harp, D.K. Keefer, and R.C. Wilson. 1985. The Borah Peak, Idaho Earthquake of
October 28, 1983–Liquefaction. Earthquake Spectra 2(4):71–89.
Youd, T.L., I.M. Idriss, R.D. Andrus, I. Arango, G. Castro, J.T. Christian, R. Dobry, W.D.L. Finn, L.F.
Harder, M.E. Hynes, K. Ishihara, J.P. Koester, S.C.C. Liao, W.F. Marcuson, III., G.R. Martin,
J.K. Mitchell, Y. Moriwaki, M.S. Power, P.K. Robertson, R.B. Seed, and K.H. Stokoe, II. 2001.
Liquefaction resistance of soils: Summary report from the 1996 NCEER and 1998 NCEER/NSF
Workshops on Evaluation of Liquefaction Resistance of Soils. Journal of Geotechnical and
Geoenvironmental Engineering 127:817–833.
Youd, T.L., C.M. Hansen, and S.F. Bartlett. 2002. Revised multilinear regression equations for prediction
of lateral spread displacement. Journal of Geotechnical and Geoenvironmental Engineering
128(2):1007–1017.
Zeghal, M., and A.-W. Elgamal. 1994. Analysis of site liquefaction using earthquake records. Journal of
Geotechnical Engineering, ASCE 120(6):996–1017.
Zeghal, M., and U. El Shamy. 2004. A continuum-discrete hydromechanical analysis of granular deposit
liquefaction. International Journal for Numerical and Analytical Methods in Geomechanics
28(14):1361–1383.
Zeghal, M., A.-W. Elgamal, and E. Parra. 1996. Analyses of site liquefaction using downhole array
seismic records. Paper no. 371 in Proceedings, 11th World Conference on Earthquake
Engineering. 8 pp.
Zeghal, M., T. Abdoun, A. Elmekati, V. Bennett, and L. Danisch. 2007. Local identification of soil and
soil-structure systems using shape-acceleration arrays. In Seventh International Symposium on
Field Measurements in Geomechanics, FMGM 2007, Boston, Massachusetts. 10 pp.
Zhang, G., P.K. Robertson, and R.W.I. Brachman. 2002. Estimating liquefaction-induced ground
settlements from CPT for level ground. Canadian Geotechnical Journal 39:1168–1180.
Zhang, G., P.K. Robertson, and R.W.I. Brachman. 2004. Estimating liquefaction-induced lateral
displacements using the standard penetration test or cone penetration test. Journal of
Geotechnical and Geoenvironmental Engineering 130(8):861–871.
Zhodi, T., and P. Wriggers. 2005. An Introduction to Computational Micromechanics. New York:
Springer.
Zhu, J., D. Daley, L.G. Baise, E.M. Thompson, D.J. Wald, and K.L. Knudsen. 2014. A geospatial
liquefaction model for rapid response and loss estimation. Earthquake Spectra 31(3):1813–1837,
doi:10.1193/121912EQS353M.
Zienkiewicz, O.C., and T. Shiomi. 1984. Dynamic behavior of saturated porous media: The generalized
Biot formulation and its numerical solution. International Journal for Numerical Analytical
Methods in Geomechanics 8(1):71–96.
REFERENCES 245

Zienkiewicz, O.C., and R.L. Taylor. 2013. The Finite Element Method, 7th Ed. New York: McGraw-Hill.
Zienkiewicz, O.C., A.H.C. Chan, M. Pastor, D.K. Paul, and T. Shiomi. 1990a. Static and dynamic
behaviour of soils: a rational approach to quantitative solutions. I. Fully saturated problems.
Proceedings of Royal Society of London A 429:285–309.
Zienkiewicz, O.C., Y.M. Xie, B.A. Schrefler, A. Ledesma, and N. Bicanic. 1990b. Static and dynamic
behaviour of soils: a rational approach to quantitative solutions. II. Semi-saturated problems.
Proceedings of Royal Society of London A 429:311–321.
Zienkiewicz, O.C., A.H.C. Chan, M. Pastor, B.A. Schrefler, and T. Shiomi. 1999. Computational
Geomechanics with Special Reference to Earthquake Engineering. Chichester, UK: John Wiley
& Sons, Inc.
Ziotopoulou, K., R.W. Boulanger, and S.L. Kramer, 2012. Site response analysis of liquefying sites. Geo-
Congress 2012: State of the Art and Practice in Geotechnical Engineering 1799–1808, doi:
10.1061/9780784412121.185.
246
Appendixes

247
Appendix A
Biographical Sketches of Committee Members

EDWARD KAVAZANJIAN, JR. (NAE) (Chair), is the Ira A. Fulton Professor of


Geotechnical Engineering in the School of Sustainable Engineering and the Built Environment at
Arizona State University in Tempe. He joined the faculty at Arizona State University in August
2004, after 20 years as a practicing geotechnical engineer. Dr. Kavazanjian’s expertise includes
geotechnical engineering for infrastructure development with a focus on seismic design. He is
particularly well known for his work on seismic design of transportation facilities and waste
containment systems and has delivered keynote addresses and state-of-the-art papers on these
topics at international conferences. Dr. Kavazanjian is the lead author of the U.S. Federal
Highway Administration guidance document LRFD Seismic Analysis and Design for
Geotechnical Transportation Features and Structural Foundations. He has served as principal or
co-principal investigator on National Science Foundation and California Integrated Waste
Management Board research projects on seismic hazard mitigation and waste containment
system design and performance, and he is also co-author of the U.S. Environmental Protection
Agency’s guidance document on seismic design for municipal solid waste landfill facilities. Dr.
Kavazanjian is a registered professional engineer in Arizona, California, and Washington State,
and a Diplomate of the Academy of Geo-Professionals. He holds B.S. and M.S. degrees in civil
engineering from Massachusetts Institute of Technology, and a Ph.D. in geotechnical
engineering from the University of California, Berkeley.

JOSÉ E. ANDRADE is a professor in the Division of Engineering and Applied Science at


California Institute of Technology, after serving 4 years at Northwestern University as an
assistant professor in Theoretical and Applied Mechanics. His research interests lie in the area of
computational mechanics with application to problems at the interface of physics and mechanics
to develop predictive analytical and numerical models for granular/porous materials (e.g., soils,
rocks, foam, bone, etc.). In the area of earthquake-induced liquefaction, Dr. Andrade has
proposed predictive models based on physical principles as an alternative to empirical methods.

248
APPENDIX A 249

Dr. Andrade serves on the editorial board of four leading international journals in the field: Acta
Geotechnica, Journal for Numerical and Analytical Methods in Geomechanics, Computers &
Geotechnics, and Computational Particle Mechanics. Dr. Andrade is the recipient of several
honors and awards including the 2006 Zienkiewicz Medal in computational mechanics, a 2010
National Science Foundation (NSF) Faculty Early Career Development Program award, the 2010
Young Investigator Award from the U.S. Air Force Office of Scientific Research (AFOSR), the
2011 Arthur Casagrande Career Development Award from the American Society of Civil
Engineers, and the 2011 Rocafuerte Medal for Scientific and Technological Advancements from
the Republic of Ecuador. His work is currently funded by AFOSR, the Defense Threat Reduction
Agency, NSF, and NASA. Dr. Andrade received his B.S. degree from Florida Institute of
Technology and his M.S. and Ph.D. degrees in geomechanics from Stanford University.

KANDIAH “ARUL” ARULMOLI is the president and principal of Earth Mechanics, Inc., a
geotechnical and earthquake engineering consulting firm specializing in transportation
infrastructure. He has more than 30 years of experience in soil-structure interaction, seismic site
response evaluation, and field and laboratory testing. He has a broad-based technical background
in geotechnical engineering, and is a recognized expert in the area of geotechnical earthquake
engineering. Since 1978, he has contributed to the improvement of design of California’s
infrastructure to withstand earthquakes through better understanding of the behavior of the
geotechnical materials upon which the structures are founded. Dr. Arulmoli was a member of the
California Seismic Safety Commission from 2005 to 2009 and from 2010 to 2011. While serving
on the Commission, he chaired the committee that reviewed the Pacific Earthquake Engineering
Center’s activities and served on other Commission committees and activities to help improve
the seismic safety of Californians. He was a member of the American Society of Civil Engineers’
Coasts, Oceans, Ports, and Rivers Institute team that visited Chile in April 2010 to access and
learn from the effects of the February 27, 2010 magnitude 8.8 Chile earthquake. Dr. Arulmoli
was responsible for evaluating performance of two sites at the Port of Los Angeles that suffered
damage during the January 1994 Northridge earthquake. The evaluation included various types
of geotechnical analysis, ranging from simplified liquefaction analysis to finite-element
computer modeling, to verify the past analysis procedures used in design of the wharf structures
and to improve them for future projects. Dr. Arulmoli earned his B.S. degree in civil engineering
from the University of Sri Lanka and his M.S. and Ph.D. degrees in civil engineering with
emphasis in geotechnical engineering from the University of California, Davis. He is a licensed
civil and geotechnical engineer in California.

BRIAN F. ATWATER (NAS) is a U.S. Geological Survey geologist affiliated with the
University of Washington. His earthquake-related research began in 1983 by using alluvial-fan
deposits to estimate recurrence intervals for coseismic growth of an anticline near Coalinga,
California. Since 1985, he has focused on earthquake and tsunami hazards at subduction zones.
At the Cascadia Subduction Zone he studied geologic evidence of land-level change, tsunamis,
and liquefaction. These studies branched into geologic calibrations in Alaska and Chile;
comparisons with written records of an Edo-period tsunami in Japan; paleotsunami research in
Indonesia and Thailand; and collaborations with scientists from Pakistan. The international
umbrellas included fellowships in Japan, USAID projects on Indian Ocean shores, UNESCO
projects on the Makran Subduction Zone, and a Fulbright Fellowship in Indonesia. In a Japanese
demonstration project that focused on paleoliquefaction, Dr. Atwater used giant sediment slices
250 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

to learn that subsurface dikes and sills are common beneath Columbia River banks where
liquefaction failed to produce surficial evidence during the 1700 Cascadia earthquake. His
current work extends to earthquake-hazard assessment in the Caribbean and service on a
California advisory board on water resources and ecosystem restoration. Dr. Atwater holds B.S.
and M.S. degrees in geology from Stanford University and a Ph.D. in geology from the
University of Delaware.

JOHN T. CHRISTIAN (NAE) retired from Stone & Webster Engineering Corporation as a vice
president and is now a consulting engineer in Burlington, Massachusetts. His primary area of
interest is geotechnical engineering. Much of his early work involved developing and applying
numerical methods such as the finite element method. He has also worked on reliability methods
for geotechnical applications, soil dynamics, and earthquake engineering on a broad range of
civil engineering projects. Dr. Christian's current interests largely focus on the use of reliability
techniques in geotechnical engineering and on earthquake engineering. Much of his work in
industry was associated with power generating facilities including, but not limited to, nuclear
power plants. Dr. Christian is also interested in the evolving procedures and standards for
undergraduate education, especially as reflected in the accreditation process. He received his
B.S., M.S., and Ph.D. degrees in civil engineering from the Massachusetts Institute of
Technology.

RUSSELL GREEN is a professor of civil and environmental engineering (geotechnical) at


Virginia Polytechnic Institute and State University in Blacksburg, Virginia. Dr. Russell's
research focuses on engineering seismology, geotechnical earthquake engineering, and soil
improvement with particular emphasis on modern liquefaction evaluations, liquefaction risk
mitigation, paleoliquefaction investigations, and post-earthquake investigations. Prior to joining
the faculty at Virginia Tech in August 2008, Russell was on the faculty of the University of
Michigan for 7 years, served as a member of the technical staff (earthquake engineer) for the
U.S. Defense Nuclear Facilities Safety Board for 6 years, and served on active duty in the U.S.
Marine Corps for 4 years (honorably discharged at the rank of Sergeant). Dr. Russell is a
registered professional engineer in the Commonwealth of Virginia. He received his B.S. degree
in civil engineering from Rensselaer Polytechnic Institute, his M.S. degree in civil engineering
(structures) from the University of Illinois at Urbana-Champaign, and his Ph.D. degree in civil
engineering (geotechnical) from Virginia Tech.

STEVEN L. KRAMER joined the geotechnical group in the University of Washington’s


Department of Civil Engineering in 1984. His primary research interests include soil
liquefaction, site response analysis, seismic slope stability, and hazard analysis. Much of his
current research work is in the area of performance-based earthquake engineering, specifically
the integration of probabilistic response analyses with probabilistic seismic hazard analyses. Dr.
Kramer has received the Presidential Young Investigator Award from the National Science
Foundation, the Arthur Casagrande Professional Development Award from the American
Society of Civil Engineers (ASCE), the Walter Huber Research Prize from ASCE, and the ASCE
Norman Medal, and he was named the 2012 Academic Engineer of the Year by the Puget Sound
Engineering Council. He is the author of the book Geotechnical Earthquake Engineering and co-
developer of the computer programs ProShake and EduShake. He was a senior research scientist
in the International Centre for Geohazards at the Norwegian Geotechnical Institute in 2003 and is
APPENDIX A 251

also a member of the faculty of the European School for Advanced Studies in the Reduction of
Seismic Risk at the University of Pavia in Italy. Kramer has served as a consultant to private
firms and government agencies on earthquake-related projects in the United States and abroad.
Dr. Kramer received his B.S., M.Eng., and Ph.D. degrees from the University of California,
Berkeley.

LELIO MEJIA is a principal engineer and vice president of URS Corporation. Dr. Mejia has
been involved with a broad range of geotechnical, earthquake, dam, and foundation engineering
projects. He has extensive experience in soil liquefaction and the use of ground treatment
methods to mitigate the effects of liquefaction. He has also conducted soil-structure interaction
analyses of hydraulic structures and power plant and harbor facilities; performed seismic risk
analyses; and developed designs for earthquake ground motions for dams, industrial facilities,
bridges, and high-rise buildings. He has researched the use of three-dimensional finite element
techniques and fully nonlinear models for the dynamic response analysis of dams and earth
structure and on the mechanisms of liquefaction failure during earthquakes. He is a Secretarial
Appointee to the Advisory Committee on Structural Safety of the Department of Veterans
Affairs Facilities and is currently vice-chair of the Governance Board of the National Network
for Earthquake Engineering Simulation. He has served as a National Science Foundation (NSF)
panelist for the Faculty Early Career Development Program and other NSF research in
Geotechnical and Geohazards Systems program. Dr. Mejia has served on technical review boards
for the U.S. Bureau of Reclamation, the U.S. Army Corps of Engineers, and the California
Department of Water Resources on various dam projects and for other owners on various
engineering projects. Dr. Mejia earned his B.S. degree in civil engineering from the Universidad
Javeriana in Bogota, Colombia, and his M.S. and Ph.D. degrees in geotechnical engineering from
the University of California, Berkeley.

JAMES K. MITCHELL (NAS and NAE) is currently University Distinguished Professor


Emeritus at Virginia Polytechnic Institute and State University and a consulting geotechnical
engineer. Prior to joining Virginia Tech in 1994, he served on the faculty at the University of
California, Berkeley, where he held the Edward G. Cahill and John R. Cahill Chair in the
Department of Civil and Environmental Engineering until the time of his retirement in 1993.
Concurrent to his tenure at Berkeley, he was chair of civil engineering from 1979 to 1984 and
research engineer in the Institute of Transportation Studies and in the Earthquake Engineering
Research Center. His primary research and consulting activities have focused on experimental
and analytical studies of soil behavior related to geotechnical problems, admixture stabilization
of soils, soil improvement and ground reinforcement, physicochemical phenomena in soils,
environmental geotechnics, time-dependent behavior of soils, in-situ measurement of soil
properties, and mitigation of ground failure risk during earthquakes. He has authored more than
375 publications, including the graduate-level text and geotechnical reference, Fundamentals of
Soil Behavior. A licensed civil engineer and geotechnical engineer in California and professional
engineer in Virginia, Dr. Mitchell has served as chair or officer for numerous national and
international organizations, including chair of the U.S. National Committee for the International
Society for Soil Mechanics and Foundation Engineering and vice president of the International
Society for Soil Mechanics and Geotechnical Engineering. He chaired the National Research
Council’s (NRC’s) Geotechnical Board as well as three NRC study committees and served as a
member of several other NRC study committees. He has received numerous awards, including
252 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

the Norman Medal, the Walter L. Huber Research Prize, the Terzaghi Lecture Award and the
Outstanding Projects and Leaders Award from the American Society of Civil Engineers, and the
NASA Medal for Exceptional Scientific Achievement. He was elected to the National Academy
of Engineering in 1976 and to the National Academy of Sciences in 1998. Dr. Mitchell received
a Bachelor of Civil Engineering degree from Rensselaer Polytechnic Institute and M.S. and
Sc.D. degrees in civil engineering from the Massachusetts Institute of Technology.

ELLEN RATHJE is the Warren S. Bellows Centennial Professor of Civil Engineering at The
University of Texas at Austin. Her research encompasses the seismic stability of earth slopes,
site response modeling, liquefaction evaluation and soil improvement, and the application of
remote sensing to geotechnical phenomena. Dr. Rathje was one of the developers of a new in situ
dynamic liquefaction test, which utilizes a large, truck-mounted hydraulic shaker to induce
liquefaction in localized zones of saturated soil. This testing technique is the first of its kind and
is expanding the tools available to study liquefaction in situ. Dr. Rathje has also been involved in
centrifuge testing to evaluate soil improvement techniques such as prefabricated vertical drains.
This research involved centrifuge testing of untreated and drain-treated slopes as well as
numerical modeling of the centrifuge tests. Her current research efforts involve the use of remote
sensing to measure deformations associated with liquefaction and lateral spreading. Dr. Rathje
has also been involved in earthquake reconnaissance efforts through her participation as co-chair
of the Geotechnical Extreme Events Reconnaissance Association. She has participated in and/or
led several reconnaissance missions (1999 Kocaeli earthquake in Turkey, 2001 Bhuj earthquake
in India, 2004 Niigata-ken Chuetsu earthquake in Japan, and 2010 Haiti earthquake) and through
this work has documented the occurrence of liquefaction during earthquakes. Dr. Rathje was a
member of the Board of Directors of the Earthquake Engineering Research Institute (EERI) from
2010 to 2013 and a member of the Scientific Earthquake Studies Advisory Committee of the
U.S. Geological Survey from 2007 to 2013. She received the Huber Research Prize from the
American Society of Civil Engineers in 2010, the Hogentogler Award for outstanding paper from
ASTM Committee D18 in 2010 (for a paper on in situ liquefaction testing), the Shamsher
Prakash Research Award in 2007, and the Shah Innovation Prize from EERI in 2006.

JAMES R. RICE (NAS and NAE) is the Mallinckrodt Professor of Engineering Sciences and
Geophysics at Harvard University, in its School of Engineering and Applied Sciences and
Department of Earth and Planetary Sciences. From 1965 to 1981, Dr. Rice was a professor at the
Division of Engineering, Brown University. He holds B.S., M.S., and Ph.D. degrees in
mechanics from Lehigh University. Dr. Rice studies phenomena relating to stressing,
deformation, flow, and fracture, which has been directed in recent years to geophysics
(seismology, glaciology, tectonophysics) and to civil and environmental engineering hydrology
and geomechanics. His seismic studies focus on the nucleation of earthquake rupture, thermo-
and hydro-mechanical weakening of fault zones during seismic slip, fracture propagation through
branched and offset fault systems, tsunami generation and propagation, and relations among
stressing, seismicity, and deformation in or near continental and subduction fault systems,
including the physics of aseismic deformation transients. His research related to hydrologic
processes, including poroelastic-plastic effects and other fluid interactions in the deformation and
failure of earth materials, has application to glacial flows, including rapid and episodic ice
motions, glacial earthquakes, and massive ice-sheet under-flooding events as natural hydraulic
fractures, and also to submarine and subaerial landslide processes.
APPENDIX A 253

YUMEI WANG is a geohazards engineer at the Oregon Department of Geology and Mineral
Industries and focuses on building resilience to future earthquakes, tsunamis, and landslides. She
advised the National Research Council on landslide hazards and earthquake resilience and to the
National Earthquake Hazards Reduction Program (NEHRP), has chaired the Oregon Seismic
Safety Policy Advisory Commission, and has taken part in post-earthquake assessments
including the 2011 Tohoku, Japan, and 2010 Maule, Chile, disasters. Ms. Wang has performed
liquefaction and lateral spreading analyses that include subsurface exploration and paleoseismic
investigations, has developed liquefaction hazard maps, and is exploring liquefaction research
needs to understand critical infrastructure risk in Oregon. Ms. Wang has been a guest on PBS
NewsHour has been interviewed by The New York Times, and has appeared in documentaries
produced by NOVA and National Geographic. Ms. Wang has served as Congressional Fellow,
sponsored by the American Association for the Advancement of Science, in the U.S. Senate in
Washington, DC, and has worked as a geotechnical consultant in California. Ms. Wang has a
B.S. degree in geological sciences from the University of California, Santa Barbara, and an M.S.
degree in civil engineering from the University of California, Berkeley.
Appendix B
Meeting Agendas and Workshop Participants

Committee on State of the Art and Practice in Earthquake Induced Soil Liquefaction Assesment
Meeting 1, November 12, 2013

Keck Center of the National Academies


500 Fifth Street NW, Washington, DC 20001

AGENDA

Tuesday, November 12, 2013


OPEN SESSION
ROOM 101
9:00 A.M. – 3:00 P.M.

9:00 Welcome, introductions, overview of the Statement of Task


Dr. Edward Kavazanjian, NAE, Committee Chair

9:20 Presentations from sponsors—What are the issues facing your organization?
Dr. Dave Gillette, Bureau of Reclamation
Dr. Thomas Weaver, U.S. Nuclear Regulatory Commission
Mr. Robert Schweinfurth, Geo-Institute/American Society of Civil Engineers
Mr. Antonio Gioiello, Port of Los Angeles

10:40 Break

10:55 Changes since the 1996/1998 Workshops


Dr. T. Les Youd, NAE, Brigham Young University (Emeritus)

11:35 Industry practitioner perspective: What are the issues facing practicing engineers?

254
APPENDIX B 255

Dr. Donald Anderson, CH2M Hill

12:05 Working Lunch

1:05 Public-sector dam safety engineer perspective


Dr. Craig Davis, Los Angeles Department of Water and Power

1:35 Policy perspective: What are the issues facing policy makers?
Mr. Charles Real, California Geological Survey

2:05 General discussion about the statement of task

3:00 Adjourn Open Session


256 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

WORKSHOP ON STATE OF THE ART AND PRACTICE IN EARTHQUAKE INDUCED LIQUEFACTION


ASSESSMENT

Memorial Union
Arizona State University
301 E. Orange Street, Tempe, Arizona

March 10-11, 2014

AGENDA
(detailed agendas for each session are found in corresponding
sections of the briefing materials)

DAY ONE: Monday, March 10, 2014

7:45 a.m. Check in; Continental breakfast

8:15 Welcome and introductory remarks


Edward Kavazanjian, chair, Committee on State of the Art and Practice in
Earthquake Induced Soil Liquefaction

SESSION 1: CASE HISTORIES/DATA COLLECTION

Session moderator: Russell Green, Committee Member

8:30 Plenary Session


Statement of Session Goals

8:35 Plenary Speakers: Thomas Holzer, U.S. Geological Survey


Peter Robertson, Gregg Drilling and Testing
 What are the essential data that must be known or obtained to interpret a liquefaction
related phenomena case history?
 What are the “critical holes” in the current liquefaction related phenomena case history
databases (i.e., what important scenarios/cases are absent or underrepresented in the
current databases)? What are the viable approaches to filling these holes?

9:05 Panel Discussion


Jonathan Stewart, University of California, Los Angeles
Misko Cubrinovski, University of Canterbury
Kenji Ishihara, Chuo University, Tokyo

9:45 Break
APPENDIX B 257

9:55 Breakout Sessions

Breakout A: Uncertainty in case history documentation


Breakout B: Numerical (mechanistic), laboratory, and physical model data
Breakout C: Field evidence
Breakout D: Characterization of seismic demand

11:10 Break

11:30 Plenary Session


Breakout session summaries

Noon Lunch

SESSION 2: TRIGGERING OF LIQUEFACTION

Session moderator: Steve Kramer, Committee Member

1:15 Plenary Session


Statement of Session Goals

1:20 Plenary Speaker: Geoffrey Martin, University of Southern California

 What are the primary deficiencies in the simplified method for evaluation
of liquefaction potential, and how can they be improved?
 What is the role of other triggering evaluation procedures in current
practice and in the future?

1:40 Panel Discussion


Ricardo Dobry, Rensselaer Polytechnic Institute
Donald Anderson, Ch2M Hill
Misko Cubrinovski, University of Canterbury
Geoff Martin, University of Southern California

2:15 Break

2:25 Breakout Sessions

Breakout A: Liquefaction susceptibility


Breakout B: Characterization of demand
Breakout C: Characterization of resistance
Breakout D: Model development

3:55 Break
4:15 Plenary Session
Breakout session summaries
258 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

4:45 Description of goals for Day 2 of workshop


5:00 Adjourn for the day

DAY TWO: Tuesday, March 11, 2014

8:15 a.m. Check In; Continental breakfast

8:25 Welcome and introductory remarks


Edward Kavazanjian, chair, Committee on State of the Art and Practice in
Earthquake Induced Soil Liquefaction

SESSION 3: CONSEQUENCES OF LIQUEFACTION

Session moderator: Lelio Mejia, Committee Member

8:25 Plenary Session


Statement of Session Goals

Plenary Session topics:


 Residual strength and stress-strain-pore pressure behavior
 Advanced analytical methods

8:30 Plenary Speakers:


Ricardo Dobry, Rensselaer Polytechnic Institute
Michael Beaty, Beaty Engineering LLC
 What are the latest developments for the prediction of post-triggering
consequences?
 What are the main deficiencies in current practice for the assessment of
residual shear strength and stress-strain-pore pressure behavior, and what
research is needed to address them?

9:00 Open discussion

9:25 Break

9:40 Breakout Session

Breakout A: Residual strength characterization


Breakout B: Analytical models
Breakout C: Simplified methods for settlement and lateral spreading
Breakout D: Earth structures and soil structure interaction
11:05 Break
APPENDIX B 259

11:30 Plenary Session


Breakout session summaries

12:00 noon Lunch

SESSION 4: ALTERNATIVES TO CURRENT PRACTICE


AND PATHS FORWARD

Session Moderator: José Andrade, Committee Member

1:00 Plenary Session


Statement of Session Goals

1:15 Plenary Presentations:

Robert Kayen, U.S. Geological Survey


Giuseppe Buscarnera, Northwestern University
Ronaldo Borja, Stanford University
 Mechanics and physics-based models for liquefaction and its
consequences and their uncertainties i.e., Multiphase models, numerical
models, analytical, constitutive models, etc.
 Alternative methods and models for liquefaction and their uncertainties;
i.e., strain-based methods, energy-based methods, other possibilities

2:25 Break

2:35 Breakout Sessions

Breakout A: Strain and energy-based approaches


Breakout B: Analytical models
Breakout C: Post-triggering flow behavior
Breakout D: Post-triggering stress-strain behavior

3:55 Break

4:15 Plenary Session


Breakout session summaries

CONCLUDING REMARKS

4:30 Closing Remarks


Edward Kavazanjian, Chair

4:45 p.m. Workshop adjourns


260 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Workshop Participants

Tarek Abdoun, Rensselaer Polytechnic Shideh Dashti, University of Colorado


Institute Boulder
Donald Anderson, CH2M Hill Richard Davidson, URS Corporation
Scott Anderson, Federal Highway Craig Davis, Los Angeles Department of
Administration Water and Power
José Andrade, California Institute of Armen Der Kiureghian, University of
Technology California, Berkeley
Pedro Arduino, University of Washington Steven Dickenson, New Albion
Kandiah “Arul” Arulmoli, Earth Mechanics, Geotechnical, Inc.
Inc. Ricardo Dobry, Rensselaer Polytechnic
Brian Atwater, U.S. Geological Institute
Survey/University of Washington Roupen Donikian, Parsons Brinckerhoff
Laurie Baise, Tufts University Elizabeth Eide, The National Academies of
Jean-Pierre Bardet, The University of Texas Sciences, Engineering, and Medicine
at Arlington Kevin Franke, Brigham Young University
Steve Bartlett, University of Utah Ian Friedland, Federal Highway
Chris Baxter, The University of Rhode Administration
Island David Gillette, Bureau of Reclamation
Michael Beaty, Beaty Engineering LLC Dan Gillins, Oregon State University
Ronaldo Borja, Stanford University Russell Green, Virginia Tech
Ross Boulanger, University of California, Les Harder, HDR Engineering
Davis Youssef Hashash, University of Illinois
Scott Brandenberg, University of California, Tom Holzer, U.S. Geological Survey
Los Angeles I.M. Idriss, University of California, Davis
Jonathan Bray, University of California, Kenji Ishihara, Chuo University
Berkeley
Edward Kavazanjian, Jr., Arizona State
Giuseppe Buscarnera, Northwestern University
University
Robert Kayen, U.S. Geological Survey
K. Onder Cetin, Middle East Technical
University Steven Kramer, University of Washington
John Christian, Consulting Engineer Bruce Kutter, University of California,
Davis
Brady Cox, The University of Texas at
Austin Harold Magistrale, FM Global
Misko Cubrinovski, University of Sammantha Magsino, The National
Canterbury Academies of Sciences, Engineering,
and Medicine
Yannis Dafalias, National Technical
University of Athens Allen Marr, Geocomp Corporation
APPENDIX B 261

Geoffrey Martin, University of Southern Ray Seed, University of California,


California Berkeley
Neven Matasovic, Geosyntec Consultants Tom Shantz, California Department of
Lelio Mejia, URS Corporation Transportation
Jorge Meneses, GEI Consultants Jonathan Stewart, University of California,
Los Angeles
James Mitchell, Virginia Tech
Kenneth Stokoe, The University of Texas at
Yoshi Moriwaki, GeoPentech, Inc.
Austin
Scott Olson, University of Illinois
Kohji Takimatsu, Tokyo Institute of
Jonathan Porter, Federal Highway Technology
Administration
Sabanayagam Thevanayagam, SUNY
Daniel Pradel, University of California, Los Buffalo
Angeles
Sjoerd VanBellagooy, Tonkin & Taylor
Ellen Rathje, The University of Texas at
Yumei Wang, Oregon Department of
Austin
Geology and Mineral Industries
James Rice, Harvard University
Rick Wentz, Wentz-Pacific
Mike Riemer, University of California,
Derek Wittwer, Bureau of Reclamation
Berkeley
Les Youd, Brigham Young University
Peter Robertson, Gregg Drilling and Testing
Zia Zafir, Kleinfelder, Inc.
Kyle Rollins, Brigham Young University
Curt Scheyhing, Group Delta Consultants
262 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Committee on State of the Art and Practice in Earthquake Induced Soil Liquefaction
Meeting 3, May 7, 2014

Keck Center of the National Academies


500 Fifth Street NW, Washington, DC 20001

AGENDA

Wednesday, May 7, 2014


OPEN SESSION
ROOM 101
8:30 A.M. – 1 P.M.

8:30 Welcome, introductions, session overview


Dr. Edward Kavazanjian, NAE, Committee Chair

8:40 Liquefaction Resistance of Aged Soil Deposits


Dr. Ron Andrus, Clemson University

9:30 Soil Age and Liquefaction Potential


Dr. Milan Pavich, U.S. Geological Survey

10:20 Break

10:40 Discussion: other issues related to liquefaction to older sediments?

11:00 Bayesian Thinking


Dr. Gregory Baecher, University of Maryland

Noon Lunch

1:00 pm Adjourn open session


APPENDIX B 263

Committee on State of the Art and Practice in Earthquake Induced Soil Liquefaction
Open Webinar, June 9, 2014
1:00 PM EDT

Please contact Courtney Gibbs (cgibbs@nas.edu) to register

AGENDA

1:00 pm Welcome, session overview


Dr. Edward Kavazanjian, NAE, Committee Chair

1:05 Modeling the Dynamics of Landslides that Liquefy


Dr. Richard, Iverson, USGS

1:35 Questions from committee members and discussion

2:00 pm Adjourn open session


264 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Committee on State of the Art and Practice in Earthquake Induced Soil Liquefaction
Meeting 4, July 15, 2014

Beckman Center of the National Academies


100 Academy Drive
Irvine, CA 92617
949.721.2200

AGENDA

Tuesday, July 15, 2014

OPEN SESSION
BOARD ROOM
9:00 AM – 10:00 AM

9:00 Dr. Pedro Arduino, University of Washington


Numerical modeling of post liquefaction effects

10:00 Open Session Adjourns, break


Appendix C
Histograms (or parameter distributions) of
Recent Liquefaction Triggering Databases

The following series of histograms are presented for various parameters in recently
compiled SPT, CPT, and Vs liquefaction triggering databases for level ground conditions (Cetin
et al., 2004; Moss et al., 2006; Kayen et al., 2013; Boulanger and Idriss, 2014). The ranges of the
various parameters tabulated are also listed in Table 3.2. However, the histograms give a better
indication of conditions/scenarios where case histories are well represented and not so well
represented. This is important in assessing the validity of empirical and semi-empirical
liquefaction evaluation procedures for a given scenario.

265
266 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES
APPENDIX C 267

FIGURE C.1 Histograms of variables in the recently compiled SPT liquefaction databases compiled by Cetin et al.
(2004) and Boulanger and Idriss (2014).
268 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES
APPENDIX C 269

FIGURE C.2. Histograms of variables in the recently compiled CPT liquefaction databases compiled by Moss et al.
(2006) and Boulanger and Idriss (2014).
270 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

FIGURE C.3 Histograms of variables in the recently compiled Vs liquefaction databases compiled by Kayen et al.
(2013) .
APPENDIX D
General Description of Performance-Based
Design

A complete performance-based prediction for liquefaction effects accounts for the


uncertainty and variability associated with each component incorporated in the analysis (Chapter
9, Figure 9.1). The uncertainty in ground motion prediction requires that liquefaction triggering,
consequences, and the associated losses be assessed for a broad spectrum of levels of shaking
(i.e., moderate levels of shaking that occur relatively frequently and strong shaking that occurs
only rarely), not just the level of shaking associated with a specific return period. Given some
level of ground motion, response models predict levels of response (e.g., deformations, bending
moments) that are uncertain; hence, there is a range of possible response levels for each ground
motion level. The potential for damage (e.g., cracking) from all levels of response is assessed by
means of a damage model, and uncertainty in the damage model leads to a range of possible
damage levels for a given response level. Finally, a loss model predicts the uncertain potential
for losses, both direct (e.g., reconstruction cost) and indirect (e.g., economic loss), associated
with all levels of damage. When predicting loss as part of a comprehensive performance
prediction, all possible levels of ground motion, response, and damage, as well as the
probabilities of each of these levels being reached, are considered in the calculations. The
techniques used to perform these calculations can be discrete or continuous, or a combination of
both.

THE DISCRETE APPROACH

The discrete approach to performance prediction divides prediction analysis into a series of
“events.” An event can be a condition or a state of nature (e.g., water table levels, the
liquefaction potential of a soil layer), or it can be a prior condition (e.g., a blocked drain), or it
can refer to a more conventional meaning of the word “event.” The range of possible parametric
values for each event is divided into a relatively small number of “bins” that collectively cover

271
272 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

the range of values. A probability of occurrence is assigned to each bin. These probabilities are
propagated through the performance evaluation by an event tree. Event trees first came to
general notice in the WASH 1400 report on nuclear power plant safety (U. S. Nuclear
Regulatory Commission, 1975) and have since been used widely for geotechnical and other
reliability studies. Baecher and Christian (2003) present a more detailed discussion of event trees
and related reliability techniques, especially as applied to geotechnical engineering.
The basic premise of the event tree is that performance prediction can be broken down into a
series of events defined by nodes (see Box D.1), and at each node, possible alternative values of
the event are represented by branches. For example, a stratum of sand may or may not liquefy for
a given ground motion, with some probability assigned to each alternative. The probabilities may
be evaluated statistically or subjectively, in which case they are interpreted as degrees of belief;
the probabilities of all branches emanating from a particular node must add up to 1.0. There can
be more than two alternatives at a node, and it is common to lump continuous events into a finite
number of “bins.” The alternatives at each node must be exhaustive, in that they cover all
possibilities, and they are mutually exclusive, in that an outcome cannot fall into more than one
alternative. After the event tree is constructed and the probabilities for each alternative at each
node have been assigned, the probability for each path through the event tree to a terminal
branch is computed as the product of the local contingent probabilities.
A simple example of an event tree for prediction of losses due to post-liquefaction
settlement is shown in Box D.1. The result of the event tree analysis is a probabilistic description
of the expected damage and losses. Event trees are especially useful for understanding the
relative contributions of each alternative to the computed damage and losses. Results of event
tree analyses need to be presented clearly: for example, through histograms of loss measure(s)
(e.g., repair costs). When the damage states are expressed in the form of economic losses, cost-
benefit principles can be applied. When noneconomic losses are included in the assessment,
more sophisticated decision strategies must be used to examine trade-offs among alternative
designs.

BOX D.1
Event Tree for Prediction of Loss Due to Settlement

A simple, hypothetical event tree showing the prediction of loss due to post-liquefaction settlement
induced by different levels of shaking is shown in Figure 1. Branches for one path to loss level are shown,
but the full event tree would have 5 × 2 × 4 × 4 × 5 = 800 terminal branches. The probability of each event
is indicated in parentheses. The repair cost probability for a given path of events is equal to the product of
all of the probabilities of events in that path. For example, the repair cost probability for 0 to 3 dollars per
square foot for the path shown below is the product of 0.24 × 0.2 × 0.15 × 0.45 × 0.55 = 0.001782, or
0.18%. The repair cost probabilities can be obtained for all 800 terminal branches and combined to
develop a histogram of the expected repair costs.
APPENDIX D 273

FIGURE 1 Hypothetical event tree for prediction of loss due to post-liquefaction settlement. PGA is peak
ground acceleration. SOURCE: Courtesy of S. Kramer.

THE CONTINUOUS APPROACH

The discrete approach is relatively simple and intuitive, but it requires a coarse
discretization of events and variables. These limitations can be addressed by treating variables as
continuous parameters and by convolving the performance prediction process with the results of
a probabilistic seismic hazard analysis.
The Pacific Earthquake Engineering Research (PEER) Center developed a modular
framework for performance-based earthquake engineering that mirrors the process illustrated in
Chapter 9, Figure 9.1. Applying the framework begins with characterization of ground motion
intensity described in terms of an intensity measure (IM). A response model is used to predict the
response of a system of interest in terms of an engineering demand parameter (EDP), a damage
model is used to predict physical damage in terms of a damage measure (DM) from the EDP, and
a loss model is used to predict loss in terms of a decision variable (DV) from the damage
measure. The quantities IM, EDP, DM, and DV can each be single, scalar quantities or a vector
of quantities.
PEER’s framework for performance-based evaluation is encapsulated in a “framing
equation” (Cornell and Krawinkler, 2000; Deierlein et al., 2003), a triple integral that integrates
over the expected values of IM, EDP, and DM to compute the annual rate of exceedance of the
DV,  (DV ) . The triple integral equation may be written as Equation D.1:

N DM N EDP N IM
 ( DV  dv)    
k 1 j 1 i 1
P[ DV  dv | dmk ]  P[ DM  dmk | edp j ]  P[ EDP  edp j | im i ]    IM (imi ) D.1

where P[a|b] describes the conditional probability of a given b,   IM is the annual probability of
occurrence of IM = imi, and NDM, NEDP, and NIM are the number of increments of DM, EDP, and
IM, respectively. This triple integral is solved numerically for most practical problems. Accuracy
increases with increasing number of increments.
The conditional probabilities can be described by fragility functions that predict (i) the
probability of exceeding the earthquake demand parameter, EDP, given the intensity measure,
274 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

IM; (ii) the probability of exceeding the damage measure, DM, given the EDP; and (iii) the
probability of exceeding the decision variable, DV, given the DM. Figure D.1 presents example
fragility curves for the evaluation of the performance of a pipeline system subject to earthquake-
induced ground displacement. In this case, PGA is used as the IM, lateral spreading displacement
as the EDP, number of pipeline breaks per km as the DM, and repair cost as the DV. By
combining the results of probabilistic response, damage, and loss models with the results of a
probabilistic seismic hazard analysis, this approach can provide a consistent and objective
evaluation of response, damage, and loss.

FIGURE D.1 Fragility curves showing (left-right) graphical representation of output of probabilistic
response, damage, and loss models. P[D>d|PGA] is the probability of exceeding a displacement level, d,
given some PGA; P[B>b|d] is the probability of exceeding some level of pipe breaks per km, b, given a
displacement, d; and P[C>c|b] is the probability of exceeding some cost, c, given a number of pipe breaks
per km, b. SOURCE: Courtesy of S. Kramer.

The quantity  (DV ) in Equation D.1 is computed for a range of DV values and presented in
the form of a hazard curve for loss that describes how often, on average, different levels of loss
can be expected to be exceeded (Figure D.2).  (DV ) is essentially equivalent to the annual
probability of a particular loss level and the reciprocal of the return period. The hypothetical risk
curve in Figure D.2 indicates a 1% annual probability (  (DV ) = 0.01) of $22 million in repair
costs and a 0.1% annual probability (  (DV ) = 0.001) of $74 million in repair costs. This format
allows an owner to treat expected annual losses as expenses in a cash flow analysis or to judge
their cumulative effect over different exposure times to make decisions on loss mitigation
(retrofit, insurance, abandonment, etc.) using cost-benefit analyses.
The PEER framework is modular—it can be formulated to produce response and damage
hazard curves—which can be helpful for different implementations of performance-based
evaluations. In modular form, Equation D.1 can be presented as:

N IM
 EDP (edp )   P[ EDP  edp | IM  imi ]  IM (imi ) D.2
i 1
N EDP
 DM (dm)   P[ DM  dm | EDP  edp ] 
j 1
j EDP (edp j ) D.3
N DM
 DV (dv)   P[ DV  dv | DM  dmk ]  DM (dmk ) D.4
k 1

Equations D.2-D.4 are used in response-, damage-, and loss-level implementations.


APPENDIX D 275

FIGURE D.2 Risk curve for repair cost resulting from lateral spreading-induced pipeline breaks over
extent of water supply system. SOURCE: Courtesy of S. Kramer.

THE HYBRID APPROACH

It is possible to treat some parts of the performance prediction process in a continuous


manner and other parts in a discrete manner. For example, it is often difficult to characterize
damage using continuous variables: the cracking of a concrete slab may be more readily
characterized in qualitative damage categories (slight, moderate, severe) rather than continuous
measures (e.g., cumulative length, crack width, and spacing), and the occurrence of a flow slide
is binary—it either happens or it does not. In such cases, a damage probability matrix (Whitman
et al., 1974) may be employed. To construct a damage probability matrix, the continuous range
of EDP response levels is discretized into a finite number of bins; each EDP bin is associated
with a damage state (e.g., negligible, slight, moderate, severe, and catastrophic); and values in
each cell represent the probability of occurrence of the damage state given the EDP value. The
matrix can be illustrated in tabular form as shown in Table D.1. The sum of each column is
unity, providing that the EDP and DM intervals are mutually exclusive and collectively
exhaustive.

TABLE D.1 Hypothetical Damage Probability Matrix for the Definition of DM|EDP Relationship
Damage EDP interval
State, DM Description edp1 edp2 edp3 edp4 edp5
dm1 Negligible 0.90 0.50 0.30 0.15 0.05
dm2 Slight 0.07 0.35 0.45 0.45 0.15
dm3 Moderate 0.02 0.10 0.15 0.25 0.20
dm4 Severe 0.01 0.04 0.07 0.10 0.30
dm5 Catastrophic 0.00 0.01 0.03 0.05 0.30
APPENDIX E
Glossary

The definitions below of terms specific to liquefaction reflect typical usage in the
geotechnical earthquake engineering. Some of the definitions include variants used in related
fields.

Aging Time-dependent changes in engineering properties of a soil. The agents of


change may include consolidation, secondary compression, weathering,
and cementation. Where resistance to liquefaction increases with time,
aging may be reversed by liquefaction or another disturbance that
effectively resets the geotechnical clock.

Amplification or deamplification of seismic waves Modification of the amplitude and phase of


seismic waves. Caused regionally by structures in Earth’s crust (e.g.,
amplification in basins that are filled with soft sediments) and locally by
near-surface deposits (e.g., deamplification by a liquefied horizon).

Arias intensity (Ia) The integral of the square of the acceleration in an accelerogram over
time, multiplied by π and divided by 2g, usually expressed in m/sec or
cm/sec.

Becker hammer Device used to measure penetration resistance in gravelly soils in situ.

Clay Soil consisting mainly of particles finer than 0.002 mm to 0.005 mm,
typically with some plasticity. The properties are determined in large
measure by the type and amount of specific clay minerals present.

276
APPENDIX E 277

Cone penetration test (CPT) An in situ soil testing procedure in which a standardized rod with
a conical tip is pushed into the soil at a constant rate. The resistance at the
tip (qc) and along a frictional sleeve (fc) are measured continuously as the
probe advances.

Consolidation Reduction of soil skeleton volume and increase in density as water flows
out of the pores and excess porewater pressure is dissipated.

Critical depth The depth in a soil profile where liquefaction is presumed most likely to
occur.

Critical layer The layer in a soil profile where liquefaction is presumed most likely to
occur.

Critical void ratio (ec) The void ratio at which soils are neither dilative nor contractive at large
shear strains.

Cross-hole wave velocity measurement A testing procedure to measure primary or secondary


wave velocities (p- and s-wave velocities, respectively) in which both the
source and the receivers for a stress wave are located at the same depth in
the soil, usually in two or more bore holes.

Cyclic resistance ratio (CRR) The capacity of a soil at a particular depth and state to resist
liquefaction triggering. It is evaluated by processing field data from
standard penetration tests (SPT), cone penetration tests (CPT), shear-wave
velocity (Vs) measurements, or other tests. These measurements are
usually calibrated, through case histories, to the minimum cyclic stress
ratios (CSRs) at which surface manifestations of liquefaction were
produced. Thus, the CRR is an estimate of the value of the CSR that can
be sustained at a particular site.

Cyclic shear strain (cyc) Shear strain induced in soil due to cyclic loading.

Cyclic stress ratio (CSR) The seismic demand induced at a particular depth in the soil, usually
expressed as the average earthquake-induced cyclic shear stress (cyc)
′ ).
divided by the initial vertical effective stress (𝜎𝑣𝑜

Damping Dissipation of energy associated with the deformations in a dynamically


loaded material or system. The critical damping ratio, commonly
shortened to damping ratio (D or β), normalizes the damping in a material
or system to the damping necessary to prevent oscillatory motion in free
vibration. A damping curve relates the damping ratio to the amplitude of
shear strain induced in the soil.

Depositional environment The setting—such as a lake, a river channel, an alluvial fan, or an


aeolian dune—in which sediments are laid down. The conditions can yield
278 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

distinctive combinations of particle size and layering (sedimentary facies)


that influence susceptibility to liquefaction (Bennett and Tinsley, 1995).

Downdrag See Negative skin friction.

Down-hole wave velocity measurement A testing procedure to measure primary or secondary


wave velocities (p- and s-wave velocities, respectively) in which the
source of a stress wave is located at the surface and the receiver is located
at depth in the soil, usually in a borehole.

Drained conditions Conditions in which the hydraulic conductivity (permeability) is so large


or loading is so slow that any excess porewater pressures dissipate and do
not contribute to the response of the soil.

Effective stress (𝝈′) The normal stress in a soil element from which the porewater pressure has
been subtracted. It represents the portion of the total stress that is
transmitted by contacts between grains in the soil skeleton, but it is not
equal to the contact stress between individual grains.

Effective stress path The path of the effective stresses during loading or unloading of a soil
sample as plotted in some specified stress space.

Effective stress principle A basic tenet of soil mechanics holding that the volume change and
strength behavior of a soil are governed by the effective stress rather than
the total stress.

Excess pore pressure Porewater pressure in excess of the value for hydrostatic or steady-state
flow conditions. The excess pore-pressure ratio (ru) is the excess
porewater pressure divided by the initial vertical effective stress.

Fines content The percentage by weight of the solid soil material consisting of particles
finer than fine sand—usually taken as less than 0.075 mm.

Flow failure Gravity-driven mass transport of soils in a fluid-like manner over


distances commonly measured in tens or hundreds of meters.

Ground failure Permanent differential ground movement capable of damaging


infrastructure.

Ground oscillation Cyclic movement of the ground due to earthquake shaking, usually felt and
observed at the surface.

Holocene An interval of geologic time spanning the past 11,700 sidereal years
(http://quaternary.stratigraphy.org/definitions).
APPENDIX E 279

Hydraulic fill Landfill emplaced by transporting sediment through pipes as a watery


suspension and depositing the slurry at the site. The resulting deposit is
commonly loose, sorted by grain size, and highly susceptible to
liquefaction.

Intensity A qualitative measure of the severity of earthquake shaking at a particular


place. Determined from observations of the earthquake’s effects on
humans, buildings, and the Earth’s surface. Compare with Magnitude,
below.

Intensity measure A quantitative measure of ground motion characteristics, such as peak


ground acceleration or Arias intensity.

Isotropic consolidation line (ICL) Plot of effective volumetric stress and strain for a test in
which the applied stress is isotropic.

Lateral spread Distributed lateral extensional movements on gentle slopes or toward a


free face. The extension is often marked by cracks at the ground surface.
The movements are presumed to involve the entire soil mass above a zone
of liquefaction. In some cases this zone may be a discrete horizon, while
in others it may be distributed through thicker sedimentary deposits.

Liquefaction The phenomena of seismic generation of excess porewater pressures and


consequent softening and loss of strength of saturated granular soils,
typically manifested by fluid-like behavior. The material is typically sand,
less commonly silt or gravel.

Liquefaction feature Geological evidence for liquefaction. May refer to features at the ground
surface (sand boils, lateral spreads), below ground (dikes, which cut
steeply across bedding; and sills, which are commonly horizontal), or
both. Sedimentologists ascribe most so-called liquefaction features to
fluidization, in which water expelled from a liquefied soil entrains
particles in its path and moves from regions of high to low hydraulic head.
In liquefaction assessment this distinction has been made chiefly in the
study of paleoliquefaction.

Liquefaction potential A measure of the tendency of a particular soil to liquefy due to a given
level of earthquake shaking.

Liquefaction susceptibility In the mapping of liquefaction potential, as defined by Youd and


Perkins (1978), the ability of a soil to liquefy irrespective of the level of
earthquake shaking.

Magnitude (M) A quantitative measure of the relative size of an earthquake, irrespective


of the observer’s location, often referred to as the moment magnitude
(Mw). The moment magnitude scale does not saturate and, therefore, is a
280 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

better measure for larger earthquakes than are other magnitude scales,
which include the local, or “Richter,” magnitude, ML. Compare with
Intensity, above.

Magnitude scaling factor (MSF) A factor that adjusts the cyclic stress ratio (CSR) for
durational effects, which correlate to earthquake magnitude. The
adjustment is necessary because the standard liquefaction curves are based
on earthquakes of M 7.5; in general, a larger earthquake lasts longer and
has more cycles of loading, while a smaller earthquake has fewer cycles.

Modulus reduction curve A curve relating the reduction of the soil stiffness, as measured by the
shear modulus to the amplitude of shear strain.

Modulus of elasticity (E) Elastic modulus relating longitudinal stress to longitudinal strain in
the absence of transverse restraint. Also called Young’s modulus.

Multichannel analysis of surface waves (MASW) A procedure in which the velocities of shear
waves are determined from surface waves (Rayleigh waves) that are
generated at a point on the ground surface and received at three or more
stations located on the surface at various distances from the source.

Negative skin friction Frictional forces acting on a pile generated by the surrounding soil
moving down relative to the pile rather than supporting it. Also called
downdrag.

One-dimensional site response analysis Mathematical analysis in which the earthquake is


assumed to be composed of vertically propagating shear waves.

Overburden correction (K) A multiplicative factor used to correct the cyclic resistance ratio
for the effects of initial confining pressure different from 1 atmosphere or
100 kPa.

Paleoliquefaction Liquefaction induced by a prehistoric (or pre-instrumental) earthquake.


Typically inferred from sand boils, dikes, and (or) sills (see Liquefaction
features).

Peak ground acceleration (PGA) The maximum horizontal acceleration amplitude in a


recorded earthquake ground motion.

Performance-based design A process in which the performance of a facility being designed is


evaluated over the entire range of possible loadings, rather than for one or
more discrete ground motion return periods or events. The approach is
intended to allow rational decisions regarding appropriate design levels,
including decisions related to the added value of increasing the seismic
resistance. Financial, construction, and operational issues are considered
over the life cycle of the facility. The approach usually involves
APPENDIX E 281

probabilistic descriptions of earthquake ground motions and the facility’s


response to them.

Phase transformation point or line Point on a stress plot (such as a p-q plot) at which a soil
shifts between contractive and dilative behavior, or a line connecting such
points obtained from tests under different initial stress states.

Piezocone A form of the cone penetration test (CPT) in which the cone is
instrumented to enable measurement of porewater pressures.

Pleistocene An interval of geologic time between 11,700 years ago and 2.58 million
years ago (http://quaternary.stratigraphy.org/definitions).

Poisson’s ratio (ν) Ratio between horizontal strain and vertical strain in a longitudinally
loaded elastic sample without transverse restraint.

Porewater pressure (u) The pressure in the fluids filling the pores of a soil. Sometimes called
the pore pressure.

Post-liquefaction settlement Settlement that occurs following soil liquefaction, usually due to
dissipation of excess porewater pressure.

Primary wave A stress wave in the body of a geologic deposit in which the particle
motion is in parallel with the direction of wave propagation. Also called a
p-wave or pressure wave.

Principle of effective stress See Effective stress principle.

Probabilistic Liquefaction Hazard Analysis (PLHA) A probabilistic estimation of the


exposure to liquefaction at a site.

Probabilistic Seismic Hazard Analysis (PSHA) An estimation of a site’s or a region’s


exposure to seismic motion in which the output and input parameters are
described probabilistically.

P-wave velocity (vp or Vp) The propagation velocity of a p-wave.

Relative density (DR) A measure of the density of a granular soil between its loosest and densest
states, as determined from standardized laboratory tests. It can be
expressed in terms of minimum and maximum void ratios instead of
densities. It is defined as (emax – e) / (emax – emin), where e denotes void
ratio.

Residual strength Shear strength in a soil that has liquefied, and which may be undergoing
shear strains orders of magnitude greater than those at which the
liquefaction was triggered.
282 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Response spectrum A plot of the maximum responses of similarly damped single degree of
freedom oscillators subjected to the same earthquake base motion. The
maximum responses are displayed as a function of the natural period or
frequency of the oscillators.

Sand A soil with particles in the size range 0.075–4.75 mm (engineering) or


0.062–2.0 mm (geology); or a granular soil dominated by such particles.

Sand boil Conical deposit of sand resulting from eruption of sand-bearing water. The
term is sometimes reserved for eruptions that are driven by seepage
beneath a levee, in which case a sand blow is associated with earthquake-
induced liquefaction.

Saturation In geotechnical engineering, a condition in which the voids are completely


filled with water. When only part of the void space is filled, the degree of
saturation is the ratio between the volume of water in the voids and the
total volume of the voids expressed as a percentage.

Seismic compression Earthquake-induced reduction of volume in a partially saturated soil.

Seismic cone A form of the cone penetration test (CPT) in which a sensor is mounted to
detect incoming stress waves.

Shear modulus (G) Elastic modulus (stiffness) relating shear stress to shear strain.

Shear strain amplitude The maximum increase or decrease in the shear strain during cyclic
loading.

Shear wave A stress wave in the body of a rock or soil in which the particle motion is
perpendicular to the direction of propagation. Also called a secondary
wave and an s-wave.

Shear wave velocity (Vs) The propagation velocity of a shear wave.

Silt A granular soil with particle sizes in the range 0.002–0.005 to 0.075 mm
(engineering) or 0.004 to 0.062 mm (geology).
Simple shear test A laboratory test in which a sample is loaded with horizontal shear
independently of the vertical and horizontal loading.

Soil Engineering term for unconsolidated earth material above bedrock. Not
limited to weathered horizons in which plants may be rooted.

Soil behavior type index (Ic) A parameter derived from the normalized tip and frictional
resistance in the cone penetration test and used to infer soil type.
APPENDIX E 283

Spectral acceleration (sa) or spectral velocity (sv) The value of the acceleration or velocity at a
particular frequency in a response spectrum.

Standard penetration test (SPT) An in situ soil testing and sampling procedure in which a
standard, thick-walled, split-spoon sampler is driven into the soil at the
bottom of a borehole by a 140 lb (623 Newton) hammer dropping a
distance of 30 in (0.76 m). The results are reported in terms of the number
of blows needed to drive the sampler 30 cm into the ground (N). The
results are corrected, as needed, by multiplying N by correction factors for
borehole diameter (CB); deviation from 60% of the free-fall energy
reaching the sampler (CE); overburden stress different from 1 atmosphere
(1 ton/sq. ft. or 100 kPa) (CN); drill rod lengths less than 10 to 30 m (CR);
and the absence of liners in a sampler (CS). The fully corrected result is
denoted (N1)60.

State parameter (ψ) The difference between the void ratio of a soil and its void ratio at the
steady-state condition for a given effective confining stress.

Static shear stress correction (K) A factor multiplied by the cyclic resistance ratio (CRR) to
account for the effects of initial shear stress on horizontal planes.

Static shear stress ratio () The ratio of the static shear stress to the initial normal effective
vertical stress.

Strain controlled Describing a test in which the loading is applied in terms of strains or
displacements.

Stress controlled Describing a test in which the loading is applied in terms of forces or
stresses.

Stress reduction coefficient (rd) The shear stress at depth z for a flexible soil column divided by
the shear stress at depth z for a rigid soil column given the same peak
ground acceleration at the surface.

Stress wave A pattern of particle movement in a geologic medium induced by sudden


displacement or loading such as by an earthquake. Also called a seismic
wave.

Surface waves Stress waves whose patterns are modulated by the presence of a free
surface. The most common are Rayleigh waves in which the particles
move in vertical retrograde elliptical paths in the direction of wave
propagation and are the basis of the multichannel analysis of surface
waves (MASW) test. Other types of surface waves include Love waves in
which the particles displace parallel to Earth’s surface in a direction
perpendicular to the propagation direction.
284 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

Threshold shear strain The cyclic shear strain at which residual excess porewater pressures are
generated.

Triaxial test A laboratory procedure in which a cylindrical sample is loaded with


independent control of the vertical and horizontal loading.

Triggering The initiation of liquefaction with no consideration about the deformation


potential or instability of the soil.

Undrained Refers to a loading pattern applied to a saturated soil that occurs so rapidly
there is not time for porewater pressures to dissipate or for pore fluids to
flow (i.e., volume remains constant).

Unit weight (γ) Weight of a sample of material divided by its volume. Subscripts are used
to indicate the unit weight of water (γw) and the total unit weight of soil
(γt).

Up-hole wave velocity measurement A testing procedure in which the source of a stress wave
is located at depth in the soil and the receiver is located at the surface.

Void ratio (e) The ratio between the volume of voids in a soil and the volume of solids.

Young’s modulus (E) Elastic modulus relating longitudinal stress to longitudinal strain in the
absence of transverse restraint. Also called the modulus of elasticity.
APPENDIX E 285

NOTATION AND SYMBOLS

α static shear stress ratio


β critical damping ratio
γ, γw, γt unit weight; subscripts indicate unit weight of water, total unit weight, etc.
γcyc cyclic shear strain
ν Poisson’s ratio
σ stress (subscripts indicate direction, such as v for vertical, and prime indicates
effective stress)
τcyc cyclic shear stress
ψ state parameter
CB correction factor for boring diameter in SPT
CE correction factor for energy ratio in SPT
CN correction factor for overburden in SPT
CPT cone penetration test
CR correction factor for rod length in SPT
CRR cyclic resistance ratio
CS correction factor for sampler liner in SPT
CSR cyclic stress ratio
D critical damping ratio
DR relative density
E Young’s modulus, modulus of elasticity
e void ratio
ec critical void ratio
F normalized frictional resistance from CPT
fc frictional sleeve resistance in cone penetration test
G shear modulus
Ia Arias intensity
Ic soil behavior index from CPT
ICL isotropic consolidation line
Kα static shear stress correction factor
Kσ overburden correction factor
286 STATE OF THE ART AND PRACTICE IN THE ASSESSMENT OF EARTHQUAKE-
INDUCED SOIL LIQUEFACTION AND ITS CONSEQUENCES

M earthquake magnitude
MASW multichannel analysis of surface waves
M moment magnitude
N number of blows to advance the SPT sampler 30 cm.
(N1)60 N value corrected to effective vertical stress of 1 atmosphere and 60% of the free-
fall energy reaching the sampler
PGA peak horizontal ground acceleration
PLHA Probabilistic Liquefaction Hazard Analysis
PSHA Probabilistic Seismic Hazard Analysis
Q normalized tip resistance from CPT
qc tip resistance in cone penetration test
rd stress reduction coefficient
ru excess pore-pressure ratio
sa spectral acceleration
SPT standard penetration test
sv spectral velocity
u porewater pressure
vp, Vp p-wave velocity
Vs shear wave velocity

Спизжено у ExLib: avxhome.in/blogs/exLib


Stole src from http://avxhome.in/blogs/exLib:
tanas.olesya (avax); Snorgared, D3pZ4i & bhgvld, Denixxx (for softarchive)
My gift to leosan (==leonadin GasGeo&BioMedLover from ru-board :-) - Lover to steal and edit
someone else's
Любителю пиздить и редактировать чужое

Vous aimerez peut-être aussi