You are on page 1of 606

Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE.

For personal use only; all rights reserved.

Geo-Congress
2023
GSP
340 Geotechnical Characterization

Selected Papers from Geo-Congress 2023


Los Angeles, California • March 26–29, 2023

Edited by
Ellen Rathje, Ph.D., P.E.
Brina M. Montoya, Ph.D., P.E.
Mark H. Wayne, Ph.D., P.E.
@seismicisolation
@seismicisolation
GEOTECHNICAL SPECIAL PUBLICATION NO. 340

GEO-CONGRESS 2023
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

GEOTECHNICAL CHARACTERIZATION
SELECTED PAPERS FROM SESSIONS OF
GEO-CONGRESS 2023

March 26–29, 2023


Los Angeles, California

SPONSORED BY
The Geo-Institute of the
American Society of Civil Engineers

EDITED BY
Ellen Rathje, Ph.D., P.E.
Brina M. Montoya, Ph.D., P.E.
Mark H. Wayne, Ph.D., P.E.

Published by the American Society of Civil Engineers

@seismicisolation
@seismicisolation
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Published by American Society of Civil Engineers


1801 Alexander Bell Drive
Reston, Virginia, 20191-4382
www.asce.org/publications | ascelibrary.org

Any statements expressed in these materials are those of the individual authors and do not necessarily
represent the views of ASCE, which takes no responsibility for any statement made herein. No reference
made in this publication to any specific method, product, process, or service constitutes or implies an
endorsement, recommendation, or warranty thereof by ASCE. The materials are for general information
only and do not represent a standard of ASCE, nor are they intended as a reference in purchase
specifications, contracts, regulations, statutes, or any other legal document. ASCE makes no
representation or warranty of any kind, whether express or implied, concerning the accuracy,
completeness, suitability, or utility of any information, apparatus, product, or process discussed in this
publication, and assumes no liability therefor. The information contained in these materials should not be
used without first securing competent advice with respect to its suitability for any general or specific
application. Anyone utilizing such information assumes all liability arising from such use, including but
not limited to infringement of any patent or patents.

ASCE and American Society of Civil Engineers—Registered in U.S. Patent and Trademark Office.

Photocopies and permissions. Permission to photocopy or reproduce material from ASCE publications
can be requested by sending an e-mail to permissions@asce.org or by locating a title in ASCE's Civil
Engineering Database (http://cedb.asce.org) or ASCE Library (http://ascelibrary.org) and using the
“Permissions” link.

Errata: Errata, if any, can be found at https://doi.org/10.1061/9780784484678

Copyright © 2023 by the American Society of Civil Engineers.


All Rights Reserved.
ISBN 978-0-7844-8467-8 (PDF)
Manufactured in the United States of America.

@seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 iii

Preface
This is the third volume of six Geotechnical Special Publications (GSPs) containing papers from
the 2023 Geo-Congress: Sustainable Infrastructure Solutions from the Ground Up, held in Los
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Angeles, California on March 26–29, 2023. These six volumes contain over 350 contributions that
span 22 topic areas of the ASCE Geo-Institute’s standing technical committee as well as 11 special
topics unique to Geo-Congress 2023.

This volume is focused on geotechnical characterization, with contributions related to engineering


geology, rock mechanics, site characterization, engineering geophysics, and soil
properties/modeling.

This publication culminates two years of effort by the technical planning committee whose focus
has been to continue the success of the previous Geo-Congress conference series. Many
individuals are responsible for the content of this volume, all of whom served in the efforts to
maintain the standard set by previous proceedings. An international call for papers and a rigorous
peer review process yielded more than 350 accepted technical papers, that were presented in a
wide range of technical sessions, in addition to invited keynote presentations. Papers were
reviewed in accordance with ASCE GSP standards. Accordingly, each paper was subjected to
technical review by two or more independent peer reviewers. Publication requires concurrence by
at least two peer reviewers.

The Editors would like to express their appreciation for having been provided the opportunity to
be a part of this Congress’ organization, their sincere thanks to the numerous session chairs and
reviewers, and we hope that these proceedings will be of use to the geotechnical engineering
community for many years to come.

The Editors,

Ellen Rathje, Ph.D., P.E., F.ASCE, University of Texas at Austin


Brina M. Montoya, Ph.D., P.E., M.ASCE, North Carolina State University
Mark H. Wayne, Ph.D., P.E., M.ASCE, Tensar International

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 iv

Acknowledgments
Thanks are due to the authors, primary reviewers, session chairs, and program committee, without
whom this publication would not be possible.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2023 GeoCongress Conference Program Committee

Conference Chair
Yoga Chandran, Ph.D., P.E., M.ASCE

Technical Program Committee


Chair, Scott Brandenberg, Ph.D., P.E., M.ASCE
Beena Ajmera, Ph.D., P.E., M.ASCE
David Albus, P.E., M.ASCE
Daniel Alzamora, P.E., M.ASCE
Ben Leshchinsky, Ph.D., P.E., M.ASCE
John McCartney, Ph.D., P.E., F.ASCE
Amy Rechenmacher, Ph.D., P.E., M.ASCE
Lisa Star, Ph.D., P.E., M.ASCE

Proceedings Editors
Ellen Rathje, Ph.D., P.E., F.ASCE, University of Texas at Austin
Brina Montoya, Ph.D., P.E., M.ASCE, North Carolina State University
Mark Wayne, P.E., M.ASCE, Tensar

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 v

Contents

Engineering Geology and Site Characterization


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Use of Geostatistical Analyses for Characterizing Mine Tailings as


Compared to Geophysics Profiles.................................................................................................1
Heidi Dacayanan and Bret N. Lingwall

Potential Applications of Hyperspectral Imaging on Weak Rock


Degradation Studies in Engineering Geology............................................................................11
Jessica K. Chiu, Lena Selen, and Friederike Koerting

A Technical Guide for Assessment, Setting Up, and Protection of


Rockbolts for Hydroelectric Facilities .......................................................................................22
Valérie Fréchette, Estelle Potvin, and Marco Quirion

Use of Standard Penetration Test (SPT) to Determine Raveling Index .................................35


Timothy Copeland, Ryan Shamet, Jinwoo An, Kyungwon Park,
and Boo Hyun Nam

Integration of Downhole Data Reduction Techniques for


Determination of VS Profiles .......................................................................................................44
Ayush Kumar and Anbazhagan Panjamani

Remote Sensing Using Satellite Derived Products to Assess


Sinkhole Occurrence....................................................................................................................52
Ronald J. Rizzo and L. Sebastian Bryson

Statistical Analysis of Undrained Strength as Linear Function of Depth ..............................62


Prince Turkson and Daniel R. VandenBerge

Evaluation of the Maximum-Likelihood Estimates of Site Fundamental


Frequencies for a Subset of KiK-Net Strong Motion Data ......................................................72
Mohammad Yazdi, Ramin Motamed, and John G. Anderson

Simplified Calibration Procedure for Total Earth Pressure Cells ..........................................81


William J. Baker III and Christopher L. Meehan

HVSR Measurements to Investigate Sinkholes and Treatment


Efforts along a Roadway .............................................................................................................94
Pourya Alidoust, Siavash Mahvelati, Joseph T. Coe, Atsuhiro Muto,
Sarah McInnes, Mia Painter, and Katherine Kubiak

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 vi

Towards Implementing SCPTu Geotechnical Design


Guidelines for the State of Illinois ............................................................................................104
Cody Arnold, Jorge Macedo, Paul Mayne, Luis Vergaray, Yumeng Zhao,
Sheng Dai, Lina Pua, Bruce Miller, and Brian Laningham

Determination of Geotechnical Properties in Intermediate


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Geo-Materials with Newly Developed In Situ Test Device.....................................................114


Young-Woo Song, Jung-Hyun Ryu, Inhyun Kim, and Choong-Ki Chung

Application of Geostatistical Sequential Simulation Methods for


Probabilistic 3D Subsoil Modeling and Uncertainty
Quantification Concept and Examples ....................................................................................125
Andreas Witty, Andrés A. Peña-Olarte, and Roberto Cudmani

Using MASW and Soil Borings to Characterize Heterogeneity of an


Existing Solid Waste Landfill ...................................................................................................133
Kevin Foye

Using Machine Learning to Predict Shear Wave Velocity.....................................................142


Longde Jin, Andrew Fuggle, Haley Roberts, and Christian P. Armstrong

Magnetic Resonance Imaging for Pore Water Mapping in Soils ..........................................153


Karam A. Jaradat, Maosen Wang, and Sherif L. Abdelaziz

Geophysical Engineering

Pavement Testing Using Nondestructive MASW Approach .................................................163


Ramdev R. Gohil, Jyant Kumar, and C. R. Parthasarathy

Reducing Mode Assignment Errors in Surface Wave Inversion for


Sites with a Very Shallow Impedance Contrast Using Love Type
Surface Waves ............................................................................................................................173
Salman Rahimi and Clinton M. Wood

Investigating Freeze-Thawing Behavior of Saline Soil Using


Electrical Resistivity Measurement ..........................................................................................183
Rui Liu, Cheng Zhu, John Schmalzel, Benjamin Barrowes, Danney Glaser,
Michele Maxson, and Wade Lein

Reliability of Shallow Bedrock Depth Determination from HVSR


Measurements in Central Missouri ..........................................................................................193
Chang Chi and Brent L. Rosenblad

Geotechnical Site Characterization with 3D Ambient Noise


Tomography: Field Data Applications.....................................................................................203
Yao Wang, Khiem T. Tran, Brady Cox, and Joseph P. Vantassel

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 vii

Numerical Investigation of Full Waveform Tomography to Identify


Anomalous Conditions and Untreated Zones in Jet Grout Columns ...................................214
Pourya Alidoust, Joseph T. Coe, Yang Yang, Alireza Kordjazi,
and Siavash Mahvelati

Surface Wave Site Characterization with MATLAB and Geopsy ........................................223


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Dennis R. Hiltunen

Geo-Acoustic Signals in Geotechnical and Foundation Engineering ....................................234


Anisha Pokhrel and Sherif L. Abdelaziz

Rock Mechanics

An Artificial Neural Network Model for Predicting Microbial-Induced


Alteration of Rock Strength ......................................................................................................243
Oladoyin Kolawole, Rayan H. Assaad, Mary C. Ngoma, and Ogochukwu Ozotta

Investigating Influence of Freeze-Thaw Cycles on Sandstone


Containing Pre-Existing Joints through Discrete Element Modeling ...................................252
Chenchen Huang, Cheng Zhu, Yifei Ma, and Shaini Aluthgun Hewage

Soil Properties and Modeling

Development of a Temperature-Controlled Direct Shear Box for


Frozen Samples ..........................................................................................................................262
Hossein Emami Ahari and Beena Ajmera

Modified Hyperbolic Model for Dynamic Properties of Peaty Organic Soils ......................273
Pengfei Wang, Tristan E. Buckreis, Scott J. Brandenberg, and Jonathan P. Stewart

Study of Undrained Shear Strength of Laponite for Use as Transparent


Clay Surrogate ...........................................................................................................................283
Abdurrahman Almikati, Jorge G. Zornberg, and Rodrigo Cesar Pierozan

Threshold Sand Content and the Behavior of Sand-Gravel Mixtures .................................293


Carmine P. Polito, Jay A. Grossman, Carter Eldridge, Kylie Krawulski,
William Reils, and Grace Shebel

Effect of Particle Size Distribution on Monotonic Shear Strength and


Stress-Dilatancy of Coarse-Grained Soils................................................................................302
Mandeep Singh Basson, Alejandro Martinez, and Jason T. DeJong

Drained Clay-Pipe Interface Resistance at Low Normal Stresses and


Elevated Temperatures .............................................................................................................313
M. Basma, R. Houhou, S. S. Najjar, and S. Sadek

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 viii

Shear and Elastic Moduli of Fine-Grained Soils: Impact of


Consolidation Pressure and Plasticity Characteristics ...........................................................323
Beena Ajmera, Binod Tiwari, and Quoc-Hung (Bob) Phan

A Study of Consolidation Tests on Dredged Soils with a Large


Moisture Content in Coastal Louisiana Using a Modified Odometer ..................................331
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Omar Shahrear Apu and Jay X. Wang

Effect of Thermal Volume Change on the Permeability of Kaolin


Clay during a Heating-Cooling Cycle ......................................................................................341
Nilufar Chowdhury and Omid Ghasemi-Fare

Influence of Time-Dependent Soil Thermal Conductivity on


Performance Assessment of Energy Foundations ...................................................................351
Arjun Sivaprasad and Prasenjit Basu

Exploring Box Fixity and Platen Texture in Large-Scale Direct


Shear Testing ..............................................................................................................................362
Nicholas A. Culbreth, Jennifer E. Nicks, Michael T. Adams,
and Thomas Gebrenegus

Examination of Cone Penetration in Non-Plastic Silt with a Direct


Cone Penetration Model ............................................................................................................373
Diane M. Moug and Adam B. Price

Creep, Relaxation, and Strain Rate Effects in Central Florida Silty Sand ..........................385
Sergio Marin and Luis G. Arboleda-Monsalve

A Comparison of Approaches for Determining the Virgin Compression


Line of Remolded Saturated Soils ............................................................................................396
Alireza Shiri, Daniel R. VandenBerge, and Yousef Shiri

Geochemical Properties and Atterberg Limits of Low Saline Sand-Clay


Mixtures ......................................................................................................................................405
Uddav Ghimire, Tejo V. Bheemasetti, Patrick Kozak, and Lisa Kunza

Effect of Grain Size of Granular Soils on Shear Wave Velocity and


Electrical Resistivity for Levee Health Monitoring ................................................................415
Brittany M. Russo, Adda Athanasopoulos-Zekkos, and Jongchan Kim

Deep Learning-Based Segmentation for Field Evaluation of


Riprap and Large-Sized Aggregates ........................................................................................424
Jiayi Luo, Haohang Huang, Issam I. A. Qamhia, John M. Hart,
and Erol Tutumluer

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 ix

Computed Tomography of Sand Subjected to Heating: Analysis of


Particle Displacements...............................................................................................................435
Yize Pan, Dawa Seo, Mark Rivers, Giuseppe Buscarnera,
and Alessandro F. Rotta Loria

Pre-Drilling Effects on Vibrations and Ground Deformations


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Caused by Impact Pile Driving .................................................................................................445


Berk Turkel, Jorge E. Orozco-Herrera, Luis G. Arboleda-Monsalve,
Boo Hyun Nam, and Larry Jones

Frost Susceptibility Evaluation of Clay and Sandy Soils .......................................................455


Mohammad Wasif Naqvi, Md. Fyaz Sadiq, Bora Cetin, Micheal Uduebor,
and John Daniels

Measurement of Volumetric Deformation, Strain Localization, and


Shear Band Characterization during Triaxial Testing Using a
Photogrammetry-Based Method ..............................................................................................466
Sara Fayek, Xiong Zhang, Xiaolong Xia, Qingqing Fu, and Jeffrey Cawlfield

Small-Strain Behavior and Stress Path Rotation Angle Effects of


Hawthorn Group Sands in Central Florida ............................................................................477
A. J. Aparicio-Ortube, Luis G. Arboleda-Monsalve, David G. Zapata-Medina,
and Larry Jones

Monotonic Behavior of Ledge Point Calcareous Sands with Increasing


Particle Crushing .......................................................................................................................488
Wenjing Cai and Cassandra J. Rutherford

Laboratory Evaluation of Small Strain Elastic Parameters of Coal


Ash from Bender Element Tests ...............................................................................................494
C. S. S. U. Srikanth, B. J. Ramaiah, and Murali Krishna

Effect of Freezing-Thawing on Preconsolidation Pressure ....................................................505


Seyed Morteza Zeinali and Sherif L. Abdelaziz

Simple Modifications to a Direct Shear Device to Perform Constant


Normal Stiffness (CNS) Tests ...................................................................................................514
C. D. P. Baxter, J. Fernandez Scarioni, A. S. Bradshaw, and A. Babaee

Assessment of US Frost Depth Maps Considering Climate Change Effects ........................523


Behrooz Daneshian and Sherif L. Abdelaziz

Experimental Investigation on Thermal and Electrical Properties of


Binary Soil Mixtures ..................................................................................................................533
Gaby Vasquez, Liang Li, and Hoyoung Seo

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 x

Temperature Effects on Residual Shear Strength of Soil ......................................................541


Aidy Ung, Seyed Morteza Zeinali, and Sherif L. Abdelaziz

Effects of Temperature on Volumetric Behavior of Soil Subjected to


Freezing-Thawing Cycles ..........................................................................................................549
Bohan Zhou, Zihao Shang, and Marcelo Sanchez
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Particle Shape Effects in 3D DEM Simulations of Angle of Repose .....................................557


C. S. Sandeep and T. M. Evans

Internal Structure and Breakage Behavior of Biogenic Carbonate


Sand Grains ................................................................................................................................565
Elieh Mohtashami, C. Guney Olgun, Chenglin Wu, and Tara Selly

An Evaluation of Incremental, Constant Rate of Strain and Constant


Pressure Ratio Consolidation Testing ......................................................................................575
Ryan Lavorati, William Marr, Kaveh Zehtab, Salim Werden, and John Christian

Numerical Analyses of a Landslide in the Sensitive Saint Adelphe Clay .............................587


Tyler J. Oathes and Ross W. Boulanger

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 1

Use of Geostatistical Analyses for Characterizing Mine Tailings as Compared to


Geophysics Profiles

Heidi Dacayanan1 and Bret N. Lingwall, Ph.D., P.E., M.ASCE2


1
Dept. of Civil and Environmental Engineering, South Dakota School of Mines and Technology.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Email: heidi.dacayanan@mines.sdsmt.edu
2
Dept. of Civil and Environmental Engineering, South Dakota School of Mines and Technology,
Rapid, SD. Email: bret.lingwall@sdsmt.edu

ABSTRACT

Although geostatistical tools have been available for over 50 years, their implementation is
uneven across geotechnical industry. One barrier to more widespread use of geostatistics in
practice is the difficulty in knowing which realizations, of many error minimizing realizations,
are representative of the actual subsurface conditions. When using a set of subsurface
explorations, such as CPT, there can be many viable realizations of conditions. For a highly
variable mine tailings pile orthogonal 2D MASW profiles showing material variability via shear
wave velocity are compared to geostatistical realizations of materials between CPT soundings.
Specifically, orthogonal 2D MASW lines were conducted to image the subsurface for soft and
stiff locations potentially indicating rubble or debris in the subsurface, while CPT were
performed for geotechnical shallow and deep foundation design. In this work the geophysics are
used to “truth” the geostatistical analyses. This paper shows that 2D geophysical profiles and
geostatistics can both be used to constrain subsurface uncertainties, drive successful foundation
design, and control construction phase obstructions.

INTRODUCTION

As large tailing impoundments around the world a closed and site re-development are
proposed, a number of impoundments with areas greater than 1,000,000 m2 will require or are
currently requiring subsurface characterization to facilitate site reuse. Mine tailings
impoundments are complex depositional environments due to the pond sedimentation nature of
the placement. These sedimentation ponds develop alluvial-lacustrine patterns wherein larger
granular particles drop out of suspension closer to the effluent pipe, while fine-grained materials
settle at the far end of ponds. Ponds are formed by earthen berms, which evolve as needed by site
operators to contain ponds. The uncontrolled nature of tailings impoundments means that organic
matter including entire trees can be entrained within the tailings mass. Geochemical and
biogeochemical reactions in the chemically active environment can result in mineral
precipitation in localized and unpredictable areas giving zones of cemented material in an
otherwise soft, wet, and liquefiable impoundment. These factors contribute to a nearly random
distribution of hard and soft, fine-grained and granular, and dense and loose materials. In
contrast to site characterization of natural soils, which follow the laws of natural geologic
deposition, mine tailings geotechnical site characterization is particularly difficult. To meet the
challenges in geotechnical site characterization thus detailed, geostatistical analyses and
advanced use of geophysical surveys are proposed.
The study area for this paper has been described by Parkhill et al. (2019) and (2022). This
site was a former copper mine tailings impoundment. Figure 1 shows the site during tailings

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 2

deposition in the 1980s. Several detailed studies have been performed to characterize the
impoundment for commercial and industrial development now that gravity de-watering and self-
weight consolidation has largely completed (as of 2015). The first step in evaluating
development possibilities at the site and mitigation that may be required was to obtain current,
accurate estimates of settlement that may occur if fills are used to level the site. To this end, test
fills and footing load tests were constructed at the site to provide insight into settlement resulting
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

from initial placement of the test fill as well as settlement that occurs when a test fill is placed,
removed, and placed again to simulate surcharge loading (pre-loading an area of the site to
stresses larger than the stresses resulting from final fill and footing loads). Details of the test fills
and footing load tests can be found in Foster et al. (2019) and Parkhill et al. (2019). With long-
term settlements varying in test fill observations over 14 years up to 25-cm, development of any
structures on shallow foundations would be difficult unless the local variability under each
proposed structure was constrained. While deep foundations would be used for every heavily
loaded structure at the site, there were concerns about vastly unequal geotechnical capacities at
different points within structures that could lead to differential settlements and angular distortion
of the structures. Also, pile supported slabs had to be considered for the softest of areas, and
these are associated with very significant premium costs. Lastly, the history of the site as a
tailings pile was accompanied with known landfilling activities that included trees and other
obstructions. These obstructions would be detrimental to deep foundation installations.
Therefore, building specific studies were needed to identify both the softest zones of tailings and
to identify obstructions.
Explorations were conducted for a previous development attempt in 2004, consisting of
limited hollow-stem auger borings, test pits, and cone penetrometer testing (CPT) to accompany
placement of large test fills. Settlement calculations conducted after these explorations
determined that static settlement would be a very significant issue on the site. Additional
explorations, consisting of more than 50 borings and CPT were conducted on the site between
2016 and 2018 to accompany a second round of test fills and a suite of full-scale footing load
tests distributed across the more than 1,000,000 m2 site. During building-specific investigations,
more than 50 more borings and CPT were performed. Dozens of 1D and 2D geophysics
soundings were performed for the site. Explorations are summarized by Foster et al. (2019) and
(2021), while geophysics are summarized by Parkhill et al. (2022).

Figure 1. Site prior to grading and capping, looking south from the air circa 1985.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 3

The tailings consist of soils ranging in size from medium sand to clay size. While clay
minerals were generally not present (i.e., bentonite, illite, or kaolinite), much of the finer tailings
material would be classified geotechnically by behavior as clay (i.e., grain sizes and flat plate-
like particles). Although some of the tailings materials may be classified as clay, they have
additional properties and behavior that are not typical for the local clays due to their mineral
origins as ground copper-bearing rock. These differences are very significant. For example, the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

tailings appear to have generated a hard, desiccated crust prior to the capping in the 1990s.
Above the tailings, the Bureau of Reclamation placed compacted tailings, used as fill, compacted
to 92% Standard Proctor, used to construct ridges and drainage valleys as part of the capping
system. Some of the clay zones are much softer than other zones with no apparent pattern.
Likewise, some of the sandy zones are much looser than others.

GEOPHYSICS DERIVED SITE VARIABILITY ASSESSMENTS

The first attempt to characterize site variability aside from lithological sections from CPT
soundings and borings (which showed highly variable conditions), was a series of 2D and
orthogonal 2D Multi-spectral Analysis of Surface Waves (MASW) surface geophysics lines in
the area of proposed structures. Seismic Surface Waves methods such as MASW, SASW
(Spectral Analysis of Surface Waves) and ReMi™ (Refraction Micro tremor) use the dispersive
characteristics of surface waves to determine the variation of the shear wave velocity (Vs) with
depth. Values are acquired by analyzing seismic surface waves generated by random sources or
by a controlled impulsive source. The signal is received and recorded by linear array of
geophones. A dispersion curve is calculated from the data that shows the phase velocity of the
surface wave as a function of frequency or wavelength. A Vs depth profile is then modeled from
the dispersion curves and the Vs profile of the shallow subsurface is reported. The MASW
inversions were performed prior to CPT soundings, and judgement of the analyst was used in
selection and implementation of the inversion parameterizations.
This contouring was used to identify the lowest Vs zones, which correspond to the softest
material (whether clayey or sandy, though almost always clayey). In these softest zones two CPT
soundings were pushed to both confirm the presence of the soft zone but to also gain penetration
resistance data to couple with Vs profiling for design. Once confirmed with the CPT, footing
load tests were performed. See Foster et al. (2019) and Parkhill et al. (2019) for more details on
the footing load testing program and Parkhill et al. (2022) for use of orthogonal MASW lines.
Figure 2 shows the use a 2D geophysics line for a single full scale footing load test, while Figure
3 illustrates how orthogonal 2D lines were used for an entire building.
The large dataset of orthogonal 2D geophysical profiles for four buildings at the site, coupled
with the 2D geophysics lines for footing load tests within the building footprints provided
sufficient data that 3D models of Vs contouring with depth could be developed. 3D Vs
contouring allows the geotechnical engineer to identify trouble spots within a single structure,
and place conventional explorations at the exact spot of the potential trouble (debris, soft clay,
hard inclusions, etc.). Details of 3D profiling from MASW surface geophysics can be found in
Park et al. (2010 and 2020) and several commercial systems are available on the market.
An example of a 3D model of Vs is seen in Figure 4, courtesy of Sage Earth Science, who
performed geophysics work for the site, as presented in Parkhill et al. (2022). As Vs is correlated
tightly with a number of key geotechnical design parameters (unit weight, undrained shear
strength, compressibility and modulus, friction angle, and pre-consolidation stress), differential

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 4

settlements and differential foundation supports can be developed reliably in an individual


structure-basis for the site. In Figure 4, green contours correspond to spatially distributed low Vs
(i.e., soft) zones, while increasing yellow and orange correspond to high Vs (i.e., stiff) zones.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Comparison of CPT and MASW testing. From Parkhill et al. (2019).

Figure 3. Methodological approach for MASW use. From Parkhill et al. (2022).

Figure 4. 3D geophysical model from orthogonal MASW lines for a single structure.
Courtesy of Sage Earth Science, LLC, as presented in Parkhill et al. (2022).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 5

GEOSTATISTICAL USE OF CPT DATA

While 3D MASW results are compelling and were found to be an effective tool in site
characterization for the subsurface in tailings, there is concern that the smoothing performed on
the geophysics results, themselves at large distance intervals, could be over-simplifying the
complexity of the subsurface. The argument is that the relatively dispersed results of MASW (on
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1 to 3-m intervals) will miss critical thin layers or small pockets of anomalies less than 1-m in
thickness. Thus, a geostatistical study on the many CPT soundings within the same study
footprints as MASW were performed. Geostatistics was chosen for this study after conventional
developments of lithology from visual examination of CPT soundings showed that complex 3D
effects were occurring in the subsurface that were difficult to visualize with conventional
orthogonal 2D hand-drawn lithological sections.
CPT data presented in previous papers for this site (Parkhill et al. 2019, etc.) were first
processed as typical for study of interbedded soils (Vreugdenhil et al. 1994). Further processing
as part of thin-layer corrections were performed using the Inverse Filtering Technique developed
by Boulanger and DeJong (2018). The Matlab scripts provided by Boulanger and DeJong (2018)
were used to correct for the transitions between thinly bedded stiff and soft layers that may be
present in the highly variable tailings materials. The inverse filter technique increases the CPT
tip resistance (qc) in granular or cemented thin layers, while decreasing tip resistance in cohesive
and non-cemented thin layers so that the stress-front influence of deeper soft or stiff layers on the
thin layer of interest on the measured tip resistance are corrected for in a rational and physics-
compatible manner. Pore pressure and skin friction are also corrected in the method. Our
experience on this study is that inverse filtering greatly increased the reliability of geostatistical
predictions between CPT soundings, especially in regard to differentiating high fines-contents
sands from high sand content fine-grained soils (i.e., transitional soils).
Once processed, geolocated and inverse filtered CPT data were input into the commercial
Surfer software package from Golden Software. Surfer is a highly vetted and validated
geostatistical analysis engine with high-grade visualization tools. By utilizing the georeferenced
locations of the many CPT in the different building areas, a suite of 2D and 3D geostatistical
realizations were produced for three different numerical techniques for the following CPT
parameters: qc, Qt, fs, Ic and Q. The first three parameters are performance (i.e. stiffness,
strength and compressibility) indices, while the last two parameters are material behavior type
indices that correspond well to conventional soil classifications.
There are a host of different techniques that can be used in geostatistical analyses. Olea
(1999) present these in detail. The student of geostatistics will be familiar with various kriging
methods, use of variograms and semi-variograms, stochastic simulations, cokriging, and cross-
validation. For this study, the following three approaches were used to develop three different
sets of realizations. Each realization could be equally valid, if the assumptions of the approach
are all equally held, which is difficult to impossible to assess for complex sites. Thus, we
consider that each of the three realizations to be theoretically valid for spatially distributed soils
at depth for relatively flat ground surface. Other theoretically valid realizations also exist, in
many combinations of sill and nugget assignments, but we limit to three for this study.
1. Realization 1 uses the ordinary point kriging method with a spherical variogram and
nugget effect.
2. Realization 2 uses the ordinary point kriging method with an exponential variogram, a
quadratic trend removal, and nugget effect

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 6

3. Realization 3 uses the inverse distance to a power gridding method with a power of 2 and
a smoothing of 0.
Comparison of the three different realizations showed that approach 3, inverse distance to a
power gridding method, resulted in what appear to be the “best” realizations based on
geotechnical experience and confirmation with conventional borings and Standard Penetration
Test (SPT) sampling. Boring logs in the area were used to confirm which realizations produced
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the most accurate predictions of what material distributions and properties exist between CPT
soundings. Figures 5 and 6 show examples of CPT-derived spatial distributions of soils for two
different areas, one on the north of the site, the other on the south. Each of the four structures in
the two study areas (three structures to the north and one to the south) is about 11,500 m2.

Figure 5. CPT derived contours of Q for north study area at the site at depth of 2m.
Figures 5a to 5C are realizations 1 to 3.

Figure 6. CPT derived contours of Q for south study area at the site at depth of 2m.
Figures 6a to 6C are realizations 1 to 3.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 7

We note that in this study, soil behavior type estimates were seen to be more reliable from Q
than from Ic. Neither approach to CPT soil behavior type was perfect in comparison to SPT
samples, with both methods failing to predict soil behavior type accurately in more than 70% of
cases. The Q method was superior in granular deposits and in differentiating silty and clayey
sands versus sandy silts and sandy clays. The Ic method showed better performance in
differentiating organic clays from inorganic clays and elastic silts. Essentially, Q was better for
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

soils with Q greater than 15, while Ic was better when Q is less than 15 (plastic fine-grained).

COMPARISON OF CPT-DERIVED SPATIAL DISTRIBUTIONS VERSUS


GEOPHYSICS-DERIVED

Once geostatistical realizations in plan and profile views for 3D conditions for each structure
were completed, comparisons between geophysics-derived 3D subsurface conditions could be
compared to the 3D subsurface conditions models derived from geostatistical analysis of CPT
data. Figures 7 and 8 present the comparisons of plan view, for two structures, at two depths. The
results are presented in Vs and qc, which are both indices of stiffness. Images are at same scales
and cover the same areas for a 1:1 comparison of stiffness predictions by both methods.

Figure 7. Comparison of geophysics to CPT for a structure on the north at depth of 2m.

The trends seen in Figures 7 and 8 show that both methods capture trends in subsurface
stiffness, but exact locations of the stiff and soft pockets in mine tailings will not be duplicated
between methods. Accumulating errors in both methods, some as simple as error in survey
location, will lead to spatial differences in results. The two comparisons in Figures 7 and 8, do

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 8

show positive indications that carefully planned and executed subsurface investigations utilizing
state of the art geophysics and high-quality CPT soundings and borings can successfully identify
the spatial locations and properties of anomalously soft and stiff zones in highly variable
geologies and mine tailings. As “unforeseen” or “changed” conditions is a major source of
litigation in foundation design and construction in our industry, we see these results as a major
step forward in reducing the risk to geotechnical engineers and foundation designers.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 8. Comparison of geophysics to CPT for a structure on the south at depth of 5m.

DISCUSSION

A critical review of different geostatistical realizations compared to borings found that the
choice of settings in geostatistical analyses by the analyst can provide disparate results, despite
equal theoretical applicability. Reference literature on different settings for variograms and
kriging can be helpful for developing candidate choices for settings. However, it will not always
be apparent which settings will result in the most valid results. Word-of-mouth recommendations
for always choosing one method over another were found here to not necessarily hold true. Thus,
care should be used whenever using geostatistical techniques on CPT data, and “truthing” with
SPT is recommended. We find that CPT data does best in geostatistical analyses when soundings
are spaced on an even grid and spacing is no more than 30m. Although closer spacings are
desirable, the economics of site investigations must be balanced with the need for accuracy. We
find that closer spacings than 10m does not necessarily result in increased return on investment

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 9

for additional soundings for large building footprints. In contrast to CPT geostatistical studies,
the broad smoothing of geophysics does miss thin layers and small pockets but is less sensitive to
the choices of the analyst. Error and uncertainty in geophysical analyses should be considered.

CONCLUSION
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

In this study, the highly variable spatial distribution of different soils and geotechnical
parameters in a tailings impoundment were studied via 3D geophysical models from MASW and
use of geostatistics on CPT soundings in the same footprint as the 3D MASW studies. The
comparisons show that while geostatistical realizations of distributed CPT data are a valid means
of determining the spatial distributions of sandy versus clayey soils, and soft versus stiff zones,
the different geostatistical techniques used can produce quite different solutions. While it is
known that interpretation of surface geophysics data and development into subsurface models
has several sources of error and is subject to several types of uncertainty, it appears that 3D
geophysics models may be the superior method unless CPT soundings are placed on an even grid
and spaced apart no more than then depth of the CPT themselves. Also, the engineer or geologist
applying geostatistical tools to CPT sounding data should perform inverse filtering as part of
thin-layer and other corrections on the CPT data to help increase the consistency of geostatistical
realizations.

ACKNOWLEDGEMENTS

The authors wish to thank the contributions of The Gardener Companies, Trent Parkhill, Matt
Moriarty, Spencer Davis, Jennifer Blomquist, Troy Covill, and Byron Foster who have been co-
authors on previous studies for this site and developed the field data.

REFERENCES

Boulanger, R. W., and DeJong, J. T. (2018). “Inverse filtering procedure to correct cone
penetration data for thin-layer and transition effects.” Proc., Cone Penetration Testing 2018,
Hicks, Pisano, and Peuchen, eds., Delft University of Technology, The Netherlands, 25-44.
Foster, B. F., Covill, T., Blonquist, J., Parkhill, S. T., and Lingwall, B. N. (2021). Testing the
surcharge remediation hypothesis for secondary compression in foundation design,
Proceedings of IFCEE 2021.
Foster, B. F., Lingwall, B. N., Parkhill, S. T., and Moriarty, M. (2019). Large Test Fills on Mine
Tailings; a First Step in Building on a Superfund Site, 2019 GeoCongress.
Olea, R. O. (1999). Geostatistics for engineers and earth scientists. Kluwer Publishers, Boston.
Park, C. B. (2020). 3D MASW−field operation, data analysis and visualization: SEG 90th Ann.
Mtng., Houston, TX, October 11-16.
Park, C. B., and Taylor, C. (2010). 3D MASW Characterization of Sinkhole: A Pilot Study at
USF Geology Park, Tampa, FL: Proceedings of the Symposium on the Application of
Geophysics to Engineering and Environmental Problems (SAGEEP 2010), Keystone,
Colorado.
Parkhill, S. T., Lingwall, B. N., Foster, B. F., and Moriarty, M. (2019). A First Step in
Building on a Mine Tailings Superfund Site Part 2: Footing Load Tests, 2019
GeoCongress, ASCE.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 10

Parkhill, S. T., Covill, W. T., and Lingwall, B. N. (2022). Innovative Use of Orthogonal 2D
Seismic Surveys in Foundation Design. Proceedings Geocongress 2022, Charlotte NC.
Vreugdenhil, R., Davis, R., and Berrill, J. (1994). “Interpretation of cone penetration results in
multilayered soils.” International Journal of Numerical and Analytical Methods in
geomechanics, 18(9), 585-599.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 11

Potential Applications of Hyperspectral Imaging on Weak Rock Degradation Studies in


Engineering Geology

Jessica K. Chiu1; Lena Selen, Ph.D.2; and Friederike Koerting, Ph.D.3


1
Norwegian Geotechnical Institute, Norway; Dept. of Geoscience and Petroleum, Norwegian
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Univ. of Science and Technology, Norway (corresponding author).


Email: jessica.ka.yi.chiu@ngi.no
2
Norwegian Geotechnical Institute, Norway. Email: lena.selen@ngi.no
3
HySpex Division, Norsk Elekro Optikk, AS, Norway. Email: friederike@neo.no

ABSTRACT

Rock material characteristics such as mineral composition, fractures, and rock fabric can be
closely related to potential stability issues of underground constructions. The composition,
durability, and potential swelling are often diagnosed by use of X-ray diffraction (XRD)
analysis; however, this method does not recognize non-crystalline rock components and textural
properties of weathered and altered rocks. Hyperspectral imaging (HSI) is an emerging technique
in lithological and mineral mapping, which utilizes the spectral reflectance signature of the target
materials and can be applied both at laboratory and field settings. A pilot study has been
conducted to explore the potential of the HSI-based approach using rock chip samples from
swelling rocks down to microscopic scale, using the shortwave infrared (SWIR, 1,000–2,500
nm). By comparing the efficiency between HSI and traditional laboratory methods including
XRD analysis and petrographic studies, this study provides insights on using new approaches to
study weak rock degradation relevant for site characterization and risk management in civil
engineering projects.

INTRODUCTION

From an engineering point of view, some of the most serious problems in underground rock
excavation are directly related to alteration of previously competent rocks that reduces their
sturdiness, durability and strength, with consequences such as disintegration and swelling
(Wahlstrøm 1973; Selen 2017). During the last decade, an increasing number of cases reported
from the industry witness instabilities and collapses, whereby degradation in terms of
weathering, swelling and slaking/disintegration of rocks are among the causes. A mismatch
between the estimated rock characteristics and the experienced behavior of the in-situ rock mass
have caused serious challenges in numerous civil engineering projects (e.g. Brattli and Broch
1995; Carter et al., 2010; Skrede, 2017; Monticelli et al. 2020).
Rocks may undergo degradation when they are exposed to environmental agents (Ulusay,
2013; Selen, 2020), primarily during long-term geological transformation processes such as
metamorphism, weathering and alteration. Weathering and alteration processes lead to the
decomposition of the constituent minerals into stable or metastable secondary mineral products,
whereby different clay minerals are common and identifiable products. The progress of
weathering highly depends on the dimensions and amount of micro- and macrofractures as they
behave as preferential pathways for chemical agents (Ündül and Tuğrul 2016). The network of
fissures is inherited directly from geological structures and consists of fractures, faults, planes of

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 12

schistosity, bedding planes and cooling fractures. Argillization is a common hydrothermal


alteration process whereby rocks are converted to clay mineral aggregates, where chlorite and
montmorillonite may replace silicate minerals and reduce a previously competent rock to an
incoherent and swelling aggregate (Wahlstrom, 1973). However, degradation may also be
consequential to short-term processes during construction works, especially for the weaker rocks
(Selen, 2020). One particular feature of some degraded rocks is their ability to swell and
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

disintegrate. The weathering state of the rock mass, its composition and fracturing degree are
features which require investigation techniques which are both feasible and effective in
diagnosing the potential degrading behavior of the rock.
Rock swelling is time-dependent and a result of interplay of mineralogy, moisture sensitivity
and weathering/alteration (Selen 2020). In addition to expansion, swelling minerals occurring as
infilling or alteration products in seams or faults have a low shear strength which may contribute
to rock falls and slides in underground openings (Palmström, 1995). These rock types can be
shales, altered or weathered basalts, igneous or metamorphic rocks, or sedimentary rocks
containing mixed rock materials.
A first-hand identification of the weathering state and future degradation potential of rocks
can be done by recognizing the presence and distribution of both swelling and non-swelling clay
minerals in addition to the extent of micro- and macrofracturing of the rock.

LABORATORY METHODS TO DETECT DEGRADATION STATE OF ROCKS

To diagnose the weathering state and degradation potential of rocks, several methods may be
used to describe their composition, heterogeneity, strength properties and response to
environmental changes. Rock composition characteristics, such as mineral distribution, grain
size, structural and textural properties, porosity, microcracks/fissures and secondary minerals, are
important compositional factors influencing the rock properties, and can be assessed with
different variants of mineralogical assessments, such as XRD, thin section analysis (microscopy)
and Scanning Electron Microscopy (SEM). These compositional analyses are typically the first
investigations steps and used to decide further investigation strategies dependent on the available
time and resources.
XRD analysis is a method used to identify and determine the mineralogical composition of
the rock samples. It is common to perform a bulk analysis of the mineralogical content, and then
treat fine-fraction powder of the material with ethylene glycol to detect swelling minerals. Every
mineral or compound has a characteristic X-ray diffraction pattern, called “fingerprint”, which
can be matched against a database of over thousands of recorded phases (Dutrow and Clark
2021). Identification and quantification of minerals are carried out by comparing relative peak
heights of the crystalline phases.
A major drawback with XRD is that the method may not fully recognize clay and clay-like
components of weathered/altered rocks as the original crystalline structure is either broken or
altered into non-crystalline intermediates, including different sub-groups of swelling clays
(Banfield and Eggleton 1990; Selen 2020). As the XRD analysis method is primarily a tool for
characterization of crystalline materials, the quantification of mineral constituents can turn out to
be misleading for non-crystalline constituents such as amorphous mineral phases and/or swelling
minerals. Further, estimation of the content of swelling clay is difficult due to masking by other
abundant minerals in the diffractogram. Moreover, if the procedure includes steps for
differentiating between different types of clay minerals, the quantification of the minerals is

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 13

highly uncertain. The XRD analysis is therefore not a proper method for describing weathered
and swelling rock composition unless it is complemented with optical methods.
Structural and textural assessment by thin section analysis can be used to get an
overview of grain sizes, porosities and fabrics which can act as controlling factors on the ingress
of water and swelling behavior of the material (Selen 2020). The rock samples are often
impregnated with blue epoxy aiming to detect open discontinuities and pores. Microscopy
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

enables a qualitative assessment of the mineralogical and textural features of the rock, which can
be compared to the quantitative estimations of mineral constituents from the XRD analysis. The
distribution of minerals and fracture characteristics can be observed, which have important
implications on the rock behavior when exposed to water. In addition, the amorphous content can
be assessed and identified based on optical properties. SEM and thin section analysis may
successfully document features not detectable by use of XRD, however, the accuracy of the
results rely on the experience of the analyst and are time- and resource consuming investigation
methods.
Why a new method is needed? It is generally accepted that the cost and time are of main
concern in any engineering project, and the accuracy of the predicted geological conditions
during the planning phase plays an important role (Panthi and Nilsen, 2007). Several of the
frequently used investigation methods to assess the composition of rocks are based on extensive
laboratory procedures, have limitations related to their accuracy, are time consuming, are
dependent on special competence of the analyst, and need complementary tests to produce
valuable results to fulfil their intention. A method which visualizes the initial degradation state of
rocks in terms of degree of microfracturing, alteration and weathering, and in addition can
advocate the engineering geologist in the field, is the main motivation for this study.

METHOD

Hyperspectral imaging is a remote sensing technique that can provide a continuous spectrum
with over 100 narrow and contiguous spectral bands allowing for the resolution of distinct,
narrow spectral absorption features (Goetz et al. 1985; Chabrillat et al. 2002). A recent review by
Krupnik and Khan (2019) demonstrates that the method can be applied to identify diverse
minerals for geoscience studies at laboratory, site, and drone-borne scales; in the SWIR
wavelength region (ca. 1000 - 2500 nm), mineral groups of carbonates and phyllosilicates
relevant to this study (mica, chlorite, and clay minerals) can be distinguished spectroscopically
using their sharp absorption feature(s) caused by certain molecular bond vibrations.
In this study, HSI on the cut surface of six flysch and two serpentine rock chip samples (ca.
37 mm x 20 mm) is carried out using a HySpex SWIR-384 camera at the HySpex laboratory in
Oslo, Norway (Figure 1). The SWIR-384 camera records spectra with a spectral resolution of ca.
5.45 nm. A microscopic lens is used to acquire the 384 pixels per line, resulting in along and
across track spatial resolution of 0.052 mm/pixel.
The principles of hyperspectral data acquisition in the lab and following correction routines
are described in detail in Rogass (2017). The workflow differs slightly and in-house software
from Norsk Elektro Optikk AS (NEO) was used to retrieve the reflectance data based on
calibrated Spectralon reference panels of 5% or 20% reflectance captured within the scene. Data
preprocessing using ENVI ® image analysis software (L3Harris-Geospatial Solutions) has been
done to filter oversaturated pixels (below 10% of reflectance) from the radiance data and remove
the pixels outside the region of interest (ROI) for each sample. The hylite python package

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 14

(Thiele et al. 2021) has been used to smooth the spectra with a savitzky-golay filter (polynomial
order 2 and window size 9 or 15) and apply a convex-hull removal (i.e. continuum removal) to
the data to enhance the absorption features for mineral classification.

VNIR
SWIR
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Lenses
Light source

Reflection panel

Translation stage

Figure 1. Setup for hyperspectral imaging of rock chip samples at the HySpex laboratory
(Modified from Henriksen 2022).

Three pixel-wise mineral classification methods are used to classify six minerals that are
considered relevant for the rock degradation phenomena, including decision trees (DTs), spectral
angle mapper (SAM), and multi range spectral feature fitting (MRSFF). Figure 2 shows the list
of relevant minerals for this study and SWIR spectra of these minerals obtained from the USGS
spectral library version 7 (Kokaly et al. 2017).
Feature modelling is achieved by applying a minimum wavelength mapping (MWL) within
the hylite toolbox. Multiple gaussian features are fitted to the pixel spectrum, determining the
minimum wavelength position of found absorption features and their depth and width (van
Ruitenbeek et al. 2014; van der Meer et al. 2018). The calculated features are sorted by depth,
highlighting the deepest calculated feature as dominant followed by the second deepest feature
etc. The number of features that are calculated for each sample pixel differ depending on the
wavelength range that is used here. Two binary DTs are constructed based on the feature
positions of the deepest and second deepest MWL. The decision trees consider two wavelength
ranges: 1200-2400 nm (DT1) and 2100-2400 nm (DT2) (Figure 2). Figure 3 illustrates the
concept of DT2. For 1200-2400 nm, MWL is performed by fitting four Gaussian curves, whereas
only two Gaussian curves are fitted for 2100 – 2400 nm. The assigned class in each pixel
represents the dominating mineral(s) in the pixel. DT classification and MWL are performed
using the hylite toolbox. Minimum depth threshold, t, based on the hull corrected reference
spectra are set to eliminate pixels with a low signal-to-noise ratio.
SAM determines the angle between sample and reference spectra in the vector space at the n-
dimensional space with dimensionality equal to the number of bands (Kruse et al. 1993). The
reference spectra used in SAM classification are taken from the USGS spectral library version 7
(Figure 2). SAM classification is performed using the ENVI software in two wavelength ranges
1200-2400 nm (SAM1) and 2100-2400nm (SAM2) (Figure 2). The determined angle is
converted to a confidence value via: confidence value = 100% × (1 – spectral angle in radians)
(He and Barton 2021). A minimum confidence threshold of 95% is used for the classification.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 15
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Hull corrected reference spectra of the studied minerals collected from the USGS
Spectral Library version 7, plotted using the hylite toolbox. USGS sample IDs are listed
next to the mineral names (*= used only for MRSFF). Grey area: 5th and 90th percentile.
DT and SAM were performed for two spectral ranges (see dotted and red background).
Arrows show the spectral region per feature (left to right shoulder of the absorption)
chosen visually for MRSFF.

Figure 3. Binary DT2 for multiclass classification using absorption features within 2100-
2400 nm spectral range. Feature 0 represents the deepest feature and so on. t = threshold.

Spectral feature fitting (SFF), first proposed by (Clark et al. 1990), computes the least-square
fit between a continuum removed image and a reference spectrum at each wavelength within a
specified range. Multi range SFF (MRSFF) is performed using the ENVI ® software to fit
multiple absorption features. These absorption feature ranges are chosen manually and are
illustrated in Figure 2. For each studied mineral one reference spectrum is chosen from the
USGS spectral library, spectral absorption feature ranges were determined visually, manually
chosen for their perceived significance in the reference spectrum. Equal weights are used for the
selected ranges. For each reference mineral and each pixel, MRSFF gives a scale value that

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 16

measures the depth of the absorption features representing the mineral abundance and a root
mean square (RMS) error from the least square fitting. Threshold values for the scale value and
RMS error are selected for each mineral class based on the distribution of these values for the
whole image. If a pixel fulfils the threshold requirements for more than one mineral, the mineral
will be assigned based on the order in the 'mineral list'.
In addition, the Minimum Noise Fraction (MNF) Rotation transform (Green et al.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1988; Boardman and Kruse, 1994) is performed using the ENVI ® software for spectral
dimension reduction. The output of the MNF Rotation transform gives indications on the degree
and distribution of spectral heterogeneity of the samples.
Fractures in the SWIR images are manually mapped, with reference to the occurrence of
epoxy infills in the rock chip samples' thin sections where applicable. Fracture maps with three
classes are generated: open fracture, closed fracture, and no fracture.

RESULTS

The results from MNF Rotation Transformation, mineral classification maps and fracture
mapping are presented in Figure 4 for a qualitative analysis. Together with existing XRD
results of each sample, the percentage distributions of the classified minerals based on pixel
count of the HSI mineral maps for each sample are plotted in Figure 5 for a quantitative
analysis.
The color differences from the MNF rotation transformation plots (Figure 4c) show that there
are spectral and likely mineralogical differences within and among samples. The non-fracture
area of Flysch 6 and Flysch 7 appear to be similar and massive, while the other samples depict
certain textures. Coarse-grained minerals also stand out from the finer-grained matrix in Flysch
9, Flysch 10 and Flysch 11.
Based on Figure 4, there are obvious deviations in the results between different mineral
classification methods and the analyzed wavelength range. MRSFF shows the most similar
mineral distribution to XRD compared to DT and SAM (Figure 5). Comparing with the XRD
results, MRSFF gives a comparable percentage of calcite, except for Flysch 11. However,
chlorite is apparently over-represented in the MRSFF results, especially for the serpentinite
samples. DT and SAM do not show any matching calcite-chlorite distribution with the XRD
results. There are also major differences in the results between both methods that can be studied
qualitatively based on Figure 4. For instance, the grain marked with a black arrow in Figure 4e is
classified as calcite in DT1 and DT2 but as chlorite in SAM1, SAM2.
For clay minerals, kaolinite and montmorillonite are negligible via DT and SAM. This
coincides with the XRD results. MRSFF indicate the presence of corrensite and montmorillonite
in samples with known swelling potential and vice versa. On the other hand, corrensite is likely
over-represented in the DT and SAM results. In particular, corrensite is much more frequently
assigned within 2100 – 2400 nm than 1200-2400 nm via DT and SAM. Up to 10% of kaolinite
are identified via MRSFF in the flysch samples but no kaolinite pixels are classified in the
serpentinite samples. This contradicts with the findings from XRD. A significant percentage of
pixels are classified as illite in the flysch samples via SAM and a smaller percentage via MRSFF,
but none or negligible are found in XRD or DT. Around 20% of kaolinite is registered using
XRD in the serpentine samples, though no pixels are classified as kaolinite from HSI data
analysis.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 17

Flys.6 Flys.7 Flys.8 Flys.9 Flys.10 Flys.11 Serp.6 Serp.8 Flys.6 Flys.7 Flys.8 Flys.9 Flys.10 Flys.11 Serp.6 Serp.8
20% reflectance 5% reflectance 20% reflectance 5% reflectance

(d) Fracture (b) Reflectance


2.0
1.5
(a) RGB

1.0
0.5
0.0
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(c) MNF

closed

open
(e) DT1

(f) DT2
(h) SAM2
(g) SAM1

Legend for (e) - (i)


(i) MRSFF

Calcite/
Unclassified Calcite-Corrensite Chlorite Corrensite
(DT2)
Kaolinite-illite
Illite Kaolinite Montmorillonite
(DT1)
Flys. = Flysch; Serp. = Serpentinite

Figure 4. (a) Color-infrared RGB representation of the rock chip samples using the SWIR
bands no. 50, 130, and 220; (b) gray-scaled map showing the highest reflectance among all
bands; (c) RGB representation of the first three dimensions from MNF Rotation
transformation; (d) interpreted fracture maps; (e - h) mineral maps from DT and SAM
classification; (i) mineral map from MRSFF.

XRD MRSFF

* Corrensite identified qualitatively.

* * * *
DT1 DT2

SAM1 SAM2

Flys.6 Flys.7 Flys.8 Flys.9 Flys.10 Flys.11 Serp.6 Serp.8 Flys.6 Flys.7 Flys.8 Flys.9 Flys.10 Flys.11 Serp.6 Serp.8
Calcite/ Calcite- Kaolinite- Montmor-
Legend: Unclassified Chlorite Corrensite Illite Kaolinite
Corrensite (DT2) illite (DT1) illonite
Flys. = Flysch; Serp. = Serpentinite

Figure 5. Comparison of percentage distribution of the studied minerals XRD (Selen 2020;
Selen et al. 2021), DT, SAM and MRSFF.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 18

Wide, open fractures can be easily traced in the flysch samples in the DT1 and SAM2 results
(Figure 4). Closed fractures in Flysch 8 and Flysch 9 can be traced in the SAM1, SAM2 and
MRSF results as a different mineralogy than the surroundings or as unknown class. Table 1
shows that the mean reflectance for closed fracture pixels is higher than that for the other fracture
classes.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Table 1. Mean reflectance of pixels classified as open, closed, or non-fractures.

Mean reflectance
White reflectance
Open fracture Closed fracture Non-fracture
20% 0.24 0.30 0.22
5% 0.21 0.57 0.42

DISCUSSION

The rock chip samples are not perfectly flat. The varying thickness of the samples might
differ illumination dependent on the angle the camera is looking at the sample.
The end members for SAM, DT and MRSFF are chosen based on the known/expected
minerals presented in the samples. Manual selection of endmembers might already exclude
relevant minerals from the classification and skew the resulting class percentages. Quartz-rich
and feldspar-group minerals do not show distinct absorption features in the SWIR range,
explaining the high percentage of pixels labelled as “unclassified".
SAM performance is limited when pixel spectra represent material mixtures instead of “pure”
mineral spectra. Similarly, SAM shows limits when different materials can only be distinguished
by features present in a small number of wavelength bands (Austin et al. 2019). Using the SWIR
range or only parts of that range for the classification might limit the SAM classification results,
as well as the use of pure spectral endmember spectra from the USGS library instead of mixed
endmember spectra from the samples themselves. Labelled regions of interest in the sample
based on geochemical verification methods would likely increase SAM mapping accuracy.
This discrepancy in mineral classification between XRD and HSI data analysis is likely due
to different scales of the analyzed samples and the limitation of the methods. Bulk analysis is
carried out using XRD whereas only 2D surfaces are scanned for acquiring HSI data. Minerals
finer than the spatial resolution of the HSI data will probably show a mixed spectral signal with
other minerals in the same pixel. Calcite and chlorite, both of which have its deepest absorption
feature around 2325-2340 nm, share rather similar SWIR spectra and may not be easily
differentiated with the proposed DTs or using SAM. Serpentinite, which is not listed in the
classification scheme, is likely present and misclassified as chlorite in the serpentinite samples in
MRSFF. Corrensite, which does not have any diagnostic features in the shorter wavelengths but
the single broad feature around 2310 nm, can easily be misclassified from noisy pixels for which
thresholds are set manually for all the HSI data analysis methods in this study.
The classified clay minerals pixels from HSI data may indicate their presence in the samples,
although not always reported in the XRD results. Clay minerals, which are typically fine and
have small abundance, may have been lost during crushing for XRD.
HSI data captured with the microscopic lens resolution can trace microfractures based on the
difference in spectral features and reflectance. The differences are probably due to different
minerals in fracture infills than the surroundings, as well as topographic effect.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 19

Further studies are recommended to systemically evaluate the reliability of the XRD and HSI
data analysis results. A combination of different methods should also be investigated to improve
the robustness of HSI mineral classification of the studied minerals. A detailed thin section
analysis of these existing samples is necessary to verify the HSI mineral classifications.
Supervised machine learning using benchmarked data from thin sections may be further
considered. Other methods, such as the USGS PRISM Material Identification and Classification
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Routine (MICA) (Kokaly 2011) may be used to calibrate with the current mineral classification
results. However, corrensite is not available in the PRISM catalogue. Other unmixing techniques
would have to be explored to differentiate corrensite. A robust mineral classification scheme can
be further applied to study the implication of spatial relationship between microfracturing and
mineralogy in rock degradation.

CONCLUSION

This paper investigates the potential of using hyperspectral imaging (HSI) for classifying
minerals and microfractures associated with rock degradation. Clay minerals are common
products of rock alteration which in combination with rock fracturing can indicate the
weathering state and future degradation potential of the rocks (Gupta and Seshagiri Rao 2000).
Three common pixel-wise HSI-based mineral classification methods are tested and compared
with tradition X-ray diffraction (XRD) results. Multi range spectral feature fitting (MRSFF)
based on selected reference spectra from the USGS spectral library gives the most comparable
distribution of calcite and chlorite to the XRD results. Corrensite and montmorillonite are also
classified in samples with known swelling potentials using MRSFF. These clay minerals are
however not reported in the XRD results. Different scales of the analyzed samples and
constraints of difference methods likely contribute to the discrepancies in mineral classification
between XRD and HSI data analysis, and among the HSI-based techniques. This study provides
insights of studying microfractures using HSI, that can detect mineralogical and topographic
variations down to micrometer scale. Thin section studies and other methods such as USGS
PRISM MICA and machine learning, and the use of site-specific spectral libraries including
mixed-mineral spectra are recommended to further investigate the reliability and potential of
using HSI data in identifying minerals associated with rock degradation.

ACKNOWLEDGMENTS

We would like to thank NTNU and Statkraft AS for providing the samples, and master
student Lisa Henriksen for assisting the laboratory work. This work is supported by NGI via
STIPINST PhD grant from the Research Council of Norway (no. 323307), Bever Control AS,
and Bane NOR.

REFERENCES

Austin, K., Choros, K., Job, R., and McAree, R. (2019). Real-time mining face grade
determination using hyperspectral imaging techniques: results of research carried out under
MRIWA project M0518 at the School of Mechanical and Mining Engineering, the University
of Queensland. Distributed by Minerals Research Institute of Western Australia, Perth WA.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 20

Banfield, J. F., and Eggleton, R. A. (1990). "Analytical Transmission Electron Microscope


Studies of Plagioclase, Muscovite, and K-Feldspar Weathering." Clays Clay Miner, 38, 77–
89.
Boardman, J. W., and Kruse, F. A. (1994). "Automated spectral analysis: a geological example
using AVIRIS data north Grapevine Mountains Nevada." Proceedings ERIM Tenth Thematic
Conference on Geologic Remote Sensing Environmental Research Institute of Michigan, pp.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1–407-1–418, 1994.
Brattli, B., and Broch, E. (1995). "Stability problems in water tunnels caused by expandable
minerals. Swelling pressure measurements and mineralogical analysis." Engineering
Geolog,y 39, Issues 3–4, p. 151-169.
Carter, T. G., Castro, S. O., Carvalho, J. L., Hattersley, D., Wood, K., Barone, F. S., Yuen, C. M.
K., and Giraldo, D. (2010). "Tunnelling Issues of Chilean Tertiary Volcaniclastic Rocks."
XIII Ciclo Di Conferenze Di Meccanica Ed Ingegneria Delle Rocce, 215–236.
Chabrillat, S., Goetz, A. F., Krosley, L., and Olsen, H. W. (2002). "Use of hyperspectral images
in the identification and mapping of expansive clay soils and the role of spatial resolution."
Remote sensing of Environment, 82(2-3), 431-445.
Clark, R. N., Gallagher, A. J., and Swayze, G. A. (1990). "Material absorption band depth
mapping of imaging spectrometer data using a complete band shape least-squares fit with
library reference spectra." In Proceedings of the second airborne visible/infrared imaging
spectrometer (AVIRIS) workshop, vol. 90, pp. 176-186. JPL Publication 90-54, 1990.
Dutrow, B. L., and Clark, C. M. (2021). X-ray Powder Diffraction (XRD).
<https://serc.carleton.edu/research_education/geochemsheets/techniques/XRD.html>(Sept.
26, 2022).
Goetz, A. F., Vane, G., Solomon, J. E., and Rock, B. N. (1985). "Imaging spectrometry for earth
remote sensing." Science, 228(4704), 1147-1153.
Green, A. A., Berman, M., Switzer, P., and Craig, M. D. (1988). "A transformation for ordering
multispectral data in terms of image quality with implications for noise removal." IEEE
Transactions on geoscience and remote sensing, 26(1), 65-74.
Gupta, A. S., and Seshagiri Rao, K. (2000). "Weathering effects on the strength and
deformational behaviour of crystalline rocks under uniaxial compression state." Engineering
Geology, 56 (3-4), 257-274.
He, J., and Barton, I. (2021). "Hyperspectral remote sensing for detecting geotechnical problems
at Ray mine." Engineering Geology, 292, 106261.
Henriksen, L. H. (2022). Mineralogical assessment of rocks of hydropower tunnels subjected to
swelling. Master thesis, NTNU.
Kogure, T., Drits, V. A., and Inoue, S. (2013). "Structure of mixed-layer corrensite-chlorite
revealed by high-resolution transmission electron microcopy (HRTEM)." American
Mineralogist, 98(7), 1253-1260.
Kokaly, R. F. (2011). "PRISM: Processing Routines in IDL for Spectroscopic Measurements
(Installation Manual and User’s Guide, Version 1.0)." In USGS Open-File Report 2011-
1155, 432 P.
Krupnik, D., and Khan, S. (2019). "Close-Range, Ground-Based Hyperspectral Imaging for
Mining Applications at Various Scales: Review and Case Studies." Earth-Science Reviews,
198 (September): 102952.
Kokaly, R. F., et al. (2017). USGS spectral library version 7 data: Us geological survey data
release. USGS: Reston, VA, USA.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 21

Kruse, F. A., Lefkoff, A. B., Boardman, J. W., Heidebrecht, K. B., Shapiro, A. T., Barloon, P. J.,
and Goetz, A. F. H. (1993). "The spectral image processing system (SIPS)—interactive
visualization and analysis of imaging spectrometer data." Remote sensing of environment, 44
(2-3), 145-163.
Monticelli, J. P., Ribeiro, R., and Futai, M. (2020). "Relationship between durability index and
uniaxial compressive strength of a gneissic rock at different weathering grades." Bulletin of
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Engineering Geology and the Environment (2020) 79:1381–1397.


Palmstrom, A. (1995). RMi – a rock mass characterization system for rock engineering
purposes. PhD thesis, University of Oslo, Norway, 400 p.
Panthi, K. K., and Nilsen, B. (2007). "Predicted versus actual rock mass conditions: A review of
four tunnel projects in Nepal Himalaya." Tunnelling and Underground Space Technology,
22(2), 173–184.
Rogass, C., Koerting, F. M., Mielke, C., Brell, M., Boesche, N. K., Bade, M., and Hohmann, C.
(2017). "Translational imaging spectroscopy for proximal sensing." Sensors, 17(8).
Selen, L. (2017). Study on material properties and testing of various rock types, development of
investigation procedure and test methodology for future projects. Master thesis, NTNU.
Selen, L. (2020). Assessment on the swelling and disintegration potential of weak and weathered
rocks in water tunnels of hydropower projects—A contribution based on use of laboratory
testing methods. Doctoral thesis, NTNU.
Selen, L., Panthi, K. K., Mørk, M. B., and Sørensen, B. E. (2021). "Compositional features and
swelling potential of two weak rock types affecting their slake durability." Geotechnics, 1(1),
172-191.
Selen, L., Panthi, K. K., and Vistnes, G. (2020). "An analysis on the slaking and disintegration
extent of weak rock mass of the water tunnels for hydropower project using modified slake
durability test." Bulletin of Engineering Geology and the Environment, 79(4), 1919-1937.
Thiele, S. T., Lorenz, S., Kirsch, M., Acosta, I. C. C., Tusa, L., Herrmann, E., Möckel, R., and
Gloaguen, R. (2021). "Multi-scale, multi-sensor data integration for automated 3-D
geological mapping." Ore Geology Reviews, 136, 104252.
Ulusay, R. (2013). "Harmonizing engineering geology with rock engineering on stability of rock
slopes." In Rock Characterisation, Modelling and Engineering Design Methods, X. Feng, J.
A. Hudson and F. Tan, eds. 11–22.
Ündül, Ö., and Tuğrul, A. (2016). "On the variations of geo-engineering properties of dunites
and diorites related to weathering." Environmental Earth Sciences, 75(19), 1-15.
van Ruitenbeek, F. J., Bakker, W. H., van der Werff, H. M., Zegers, T. E., Oosthoek, J. H.,
Omer, Z. A., Marsh, S. H., and van der Meer, F. D. (2014). "Mapping the wavelength
position of deepest absorption features to explore mineral diversity in hyperspectral images."
Planetary and Space Science, 101, 108-117.
van der Meer, F., Kopačková, V., Koucká, L., van der Werff, H. M., van Ruitenbeek, F. J., and
Bakker, W. H. (2018). "Wavelength feature mapping as a proxy to mineral chemistry for
investigating geologic systems: An example from the Rodalquilar epithermal system." Int. J.
Appl. Earth Obs. Geoinf., 64, 237-248.
Wahlstrom, E. E. (1973). Tunneling in Rock (1st ed., Vol. 3). Elsevier.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 22

A Technical Guide for Assessment, Setting Up, and Protection of Rockbolts for
Hydroelectric Facilities
Valérie Fréchette1; Estelle Potvin2; and Marco Quirion3
1
Hydro-Québec, Dams and Control Structures Expertise Dept., Expertise Dam Safety and
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Infrastructure, Montréal, Canada. Email: frechette.valerie@hydro.qc.ca


2
Hydro-Québec, Performance, Evolution and Asset Characterization Dept., Research and
Innovation Production Lionel Boulet Pavilion, Varennes, QC, Canada.
Email: potvin.estelle@hydro.qc.ca
3
Hydro-Québec, Dams and Control Structures Expertise Dept., Expertise Dam Safety and
Infrastructure, Montréal, Canada. Email: quirion.marco@hydro.qc.ca
ABSTRACT
Hydro-Quebec owns more than 680 dams and dykes, and a large number of appurtenant
structures such as powerhouses, spillways, and water intakes distributed among 100
hydroelectric installations. Such large structures require deep surface and underground rock
excavations coupled with rock reinforcement systems that insure stability and safety. The
geological context for most of these hydroelectric installations requires ground support design,
which mainly relies on permanent mechanically anchored rock bolts, fully grouted after
installation. Over the years visual inspections revealed noteworthy corrosion on visible parts of
rock bolts in delimited areas. The purpose of this paper is to present the results of an industrial
research project initiated to investigate the extent of corrosion, its effect on bolt serviceability, and
possible causes of premature rock bolts aging in a hydroelectric setting. This paper will present the
importance and benefits of rock bolt environmental characterization, selection of suitable corrosion
protection, and implementation of selected protection. The outcomes of this research served for the
development of a technical guide to be used for inspection and maintenance.
INTRODUCTION
Ground support systems in existing facilities, mainly composed of permanent mechanically
anchored rockbolts which are fully grouted after installation, must resist over time to ensure their
intended function throughout their lifetime of serviceability. Rock walls of these excavations
have to be kept stable and safe with the installation of several thousand consolidation bolts in
addition to large quantities of protective mesh and pins to secure them. Stabilization of rock
faces and their protection against falling rocks are essential for the safety of workers, the
protection of equipment and the stability of rock excavations.
Hydro-Quebec owns more than 680 dams and dykes, and a large number of appurtenant
structures such as powerhouses, spillways and water intakes distributed among 100 hydroelectric
installations. Many of these structures are underground powerhouses with penstocks or headrace
tunnel, as well as surface powerhouses all excavated in rock. These surface and underground
stations, and their appurtenant structures, were excavated in the Canadian shield1 and mainly
reinforced with fully grouted mechanically anchored rock bolts.

______________________
1
The Canadian Shield is a large area of exposed Precambrian igneous or metamorphic rock. It covers
most of eastern and central Canada, and extends from the Great Lakes to the Artic Ocean, thus
surrounding Hudson Bay. It covers most of the province of Québec.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 23

Some years after the construction of a few facilities, visual inspections revealed noteworthy
corrosion on visible parts of rock bolts (and mesh pins) in delimited areas. This led to inquire
about the long-term effects of corrosion, specifically rock bolts strength, and ways to mitigate
this problem.
An industrial research project was initiated to investigate the extent of corrosion, the effect
on the serviceability and the possible causes of premature aging of rock bolts, in a hydroelectric
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

context. The goals of this project were:


1. to establish the state and strength of existing rock bolts affected by visible signs of
corrosion or other types of aging processes ;
2. to find high-performance solutions to hinder the processes of corrosion and aging of
altered rock bolts which appear unable to meet their support requirements ;
3. to write and implement a rock bolting guide that includes selection criteria for rock bolt
types, proper installation and protections that consider environmental characterization.
A comprehensive literature review of prevailing international practices and standards allowed
us to understand current methods of bolt protection from aggressive environmental factors. Most
research focused on the determination of protection and its use in a mining context (Villaescusa
et al 2008; Hassell 2008; Satola et Aroma 2004; Robinson and Tyler 1999; Baxter 1997). For
hydroelectric facilities, a rockbolt’s service life is higher than what is considered in designing
such systems in the mining industry. The appropriate level of protection proposed must be in
alignment with the facility’s estimated life expectancy.
This paper’s focus will be on the second part of this project, thus bolt protection to hinder
corrosion processes. First, corrosion categories and their effect on different parts of the rock bolt
will be addressed. Afterwards, a presentation of corrosion characterization processes will be
followed by environmental aggressiveness classification. Finally, a section, including decisions
trees, will orient corrosion protection selection, and a detailed section on protective paint choice
is presented. This paper only concerns corrosion protection applied to rockbolts and excludes
other parts of rock support systems.

CORROSION CATEGORIES AND HOW THEY AFFECT SUPPORT SYSTEMS.

Corrosion is a natural phenomenon that causes metals, such as steel and iron, to oxidize to a
lower energy state and form rust. A material’s corrosion behavior is affected by the environment
in which it is found and the characteristics of the exposed material. Various factors in
underground excavations are used to characterize the aggressiveness of the environment in which
support systems are used. Support systems are generally made of steel and used to support and
secure excavation walls.
A carbon steel reinforcing bolt can exhibit three known behaviors specific to corrosion
when exposed to a solution in a given environment, 1) Immune: no reaction between the metal
and the solution; 2) Active: the metal corrodes, dissolves in the solution, and loses mass; and 3)
Passive: the metal corrodes and produces an insoluble oxide film that covers and protects the
metal. This film slows or stops corrosion. If it is broken or dissolved, corrosion resumes
(Dorion, 2013).
The type and extent of corrosion found on the support system depends on the aggressiveness
of the environment in which it is located. Table 1 summarizes the phenomena involved, the wall
support equipment affected and the environments conducive to different corrosion categories
which affect four areas of the reinforcing bolt. Figure 1 shows these areas.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 24

Table 1. Corrosion categories that affect rock wall reinforcing support systems

Corrosion Phenomena Equipment Environments


categories
Generally More or less regular loss of material Bolt head, pin head, mesh, Acid, neutral or
aqueous over the entire surface with progressive rod that is uncovered or in weakly alkaline
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

thinning of the material. the presence of degraded environments


cement or epoxy grout
Atmospheric Metal reaction with oxygen, moisture Bolt head, pin head, mesh Moisture or
and pollutants forms a surface damp
electrolyte of non-compact oxide films environments,
that do not stop corrosion. pollutants
Microbial Related to microorganisms present in Bolt head, pin head, mesh, Presence of
the corrosion system. exposed rod bacteria
Pitting Produced by certain anions on metals Bolt and pin head, mesh, Neutral or
protected by a thin oxide film. pin or bolt shank with or alkaline water
Typically induces cavities a few tens of without cement or epoxy with chloride or
a micrometer in diameter. grout sulfate ions
Galvanic Due to the formation of an Contact between steel bolt Atmospheric,
electrochemical cell between two or pin head and galvanized aqueous
metals, or between two parts of a steel mesh
structure in contact with a different
chemical environment.
Stress driven Stress corrosion cracking, from joint Stressed rod with degraded Aqueous
action of a mechanical stress and an cement or epoxy grout environments
electrochemical reaction.

Figure 1. Types of corrosion affecting different zones of a reinforcing bolt

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 25

Atmospheric corrosion of bolt head (zone 1)

Atmospheric corrosion of carbon steel refers to the reaction of the metal with oxygen in the
air when high humidity and pollutants form an electrolyte on the surface. Oxide films and other
corrosion products formed are generally non-compact and their presence on a metal surface does
not halt corrosion. Atmospheric corrosion rate of iron depends on partial anodic and cathodic
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

reaction rates at interfaces of metal-electrolyte and oxide-electrolyte. Atmospheric corrosion


mechanism resembles that of corrosion in a liquid environment, with two differences: no
corrosion products are evacuated, and, electrolyte evaporates during drying periods, then re-
forms during high humidity periods. Atmospheric environment corrosiveness essentially depends
on three parameters: air relative humidity, temperature as well as sulfates and chloride ion
content.

Generalized aqueous corrosion – influence of seepage water pH (zones 1 to 4)

Corrosion mechanisms of carbon steel depend on water aggressiveness. The following types
of water can cause steel corrosion: acidic water, fresh water, oxygenated water and water
containing chloride, sulfate and nitrate ions. The potential-pH diagram, also called the Pourbaix
diagram, is a plot of possible thermodynamically stable phases of an aqueous electrochemical
system. It indicates thermodynamic corrosion potential of iron under different potential and pH
conditions (Pourbaix et al., 1974). It can be used to identify the immunity region of iron, the
region of passivity due to formation of solid protective oxides as Fe2O3 and Fe3O4, as well as the
region of corrosion to soluble ionic species as Fe2+ and Fe3+. Figure 2a also shows the stability
range of water. The area between the dashed lines corresponds to the range where water is
thermodynamically stable at ambient pressure. For each of the pH ranges shown in the red, blue
and green regions, corrosion of the iron will occur according to different mechanisms depending
on the potential applied to the iron electrode. Figures 2b) and c) show the areas where different
corrosion mechanisms occur on the reinforcing bolt injected with cement b) or with epoxy c).

Figure 2a) Pourbaix diagram (potential vs. pH) of iron-water system; b) oxidation
reactions of bolt steel in cement grout according to affected areas; c) oxidation reactions of
bolt steel sealed in epoxy resin according to affected areas

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 26

We have characterized the composition of seepage water from two power plants to establish
their corrosivity. There are two types of water sampled. First, more or less soft water, with a
conductivity of less than 500 µS/cm, whose pH varies between 6 and 10.1, and whose content of
cations and aggressive ions is low (sulphate and chloride). This type of water has moderate
corrosivity. The portrait of the water composition confirms that there is a link between the
aggressiveness of the water and the visual observations of the state of corrosion of the bolt heads.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Secondly, several sectors have seepage water that is very corrosive for steel and concrete.
They are characterized by acidic pH varying from 3.4 to 4.5, and a high conductivity ranging
from 1400 to 2700 µS/cm, which can be attributed to the high concentration of sulphates (700-
2000 ppm) and Ca, Mg, Al, Fe, Na, K. The high sulphate content in the presence of acidic water
must be considered with regard to the corrosion of grout, shotcrete and steel
For steel in air saturated water, theoretical corrosion current density in stagnant aerated water
is estimated at i = 38.6 x10-6 A/cm². This current value is equivalent to a corrosion rate for steel
of 0.45 mm/year (Revie, 2011). In an acidic environment (pH < 5), unprotected steel corrodes
easily, and corrosion rate increases significantly as the pH decreases. Iron corrodes and forms a
dissolved species in solution. Steel in an acidic environment is never passivated: a reaction with
dissolved oxygen occurs, but it is negligible compared solvated proton's reaction. Average steel
corrosion rate in acidic water is greater than 0.45 mm/year (Revie, 2011).

Corrosion mechanism of steel in neutral or alkaline media (10 ≥ pH ≥ 5)

In weakly acidic, neutral or weakly alkaline media, iron corrosion products are poorly
soluble and precipitate on metal surfaces. When loose or porous surface films are formed in a
wet or liquid environment, they are called passive films, which slow down corrosion reaction and
represent the range of passivation. Steel corrosion in natural water at neutral pH proceeds by
dissolved oxygen reaction (Landolt, 1993).
When a corrosive environment in this pH range contains only a low concentration of oxidants
such as dissolved oxygen, oxidant transport often limits the rate of corrosion. At pH values
below 8, this increase in potential is insufficient to promote passivation of the iron, and oxygen
increases the rate of iron corrosion. At pH values above 8, the presence of oxygen promotes
insoluble oxide film formation, probably in the form of γ-Fe2O3 (Pourbaix, 1974) which, in a
solution containing no aggressive ions, is generally protective. The average steel corrosion rate
in stagnant air-saturated fresh water in the presence of a poorly protective barrier film will be
less than 0.45 mm/year and may decrease to about 0.1 mm/year.

Corrosion mechanism of steel in a cement grout in a very alkaline environment (13.5 > pH
> 10)

In practice, carbon steel in an alkaline environment, such as in a cement grout, is very


resistant to corrosion since it is in the passivation range. In this environment, iron oxides and
oxyhydroxides are thermodynamically stable iron compounds (Bertolini, 2004). They form a
highly protective barrier film and prevent anodic dissolution of iron. Average steel corrosion rate
in a healthy cement grout will be as low as 0.001 mm/year.
To protect the bolt from the corrosive environment, surrounding water must not come into
contact with the bolt. Cement grouting increases shear strength between rock parts and protects
from corrosion. When injection is used for corrosion protection, it is well known that cement is

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 27

permeable to gases and ions. Cement is therefore a chemically active environment and certain
precautions must be taken to avoid corrosion of steel (Cigna et al., 2003). Grout usually slows
down corrosion of these bolts, but the longer the service life, the greater the risk of corrosion.
There are two main phenomena that can lead to corrosion of steel in grout: Depassivation of
steel due to degradation of cement grout by carbonation and pitting corrosion of steel. The
environments in which rock bolts are installed can be more or less aggressive. Acidic waters that
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

attack the lime of cements, especially waters containing dissolved carbon dioxide and humic
acids (peat bogs, etc.). Waters with a high sulfate content combines with tricalcium aluminate in
cement to form Candlot’s salt, which disorganizes the mortar as it swells. These sulfate-rich
waters include selenite waters and magnesian waters.
Depassivation of steel occurs when there is a pH reduction caused by CO 2 penetration from
the air into the grout. CO2 dissolves at high pH as carbonates, and precipitates as calcium
carbonate, causing local pH lowering. The oxide film no longer provides as effective a barrier
against corrosion. A grout is susceptible to carbonation if it is not properly cured, if SiO 2
particles are too large, or if additives that consume Ca(OH)2, are used, as in the case of fly ash.
Consequently, the average corrosion rate of a depassivated steel in a damaged cement grout will
be lower than 0.01 mm/year.
Penetration of Cl- ions into the grout can lead to pitting corrosion of the steel which is
protected by an oxide film. Pitting mechanisms depend on chemical nature and microstructure of
the metal, surface condition, including the presence of inclusions, defects such as scratches,
cracks or fissures in the metal, chemical composition of the electrolyte, including the
concentration of aggressive and nonaggressive anions, and temperature.

Stress corrosion (zone 4)

Stress corrosion cracking is a phenomenon that affects metal parts under stress in a corrosive
environment. Mechanical stress and aggressive environment combination can result in the
sudden failure of metal parts at a stress below the material’s expected strength. This phenomenon
has been reported in components like mine anchors (Gamboa & Atrens, 2003). Stress corrosion
cracking of steel starts and progresses slowly. It consists of three phases: initiation, propagation,
and rapid and brittle failure. The initiation phase can reportedly last months or even years before
visual inspection reveals the problem. Small initial cracks can be caused by surface defects (e.g.,
scratches), inconsistencies in the material, inclusions, etc. However, pitting corrosion is very
often the cause of cracks in reinforcing bolts, according to an Australian mining study (Crosky et
al., 2002).

Microbial corrosion or biocorrosion (zones 1 and 2)

Microbial corrosion concerns material which corrodes more rapidly under the effect of
microorganisms present on its surface. Biocorrosion is not a new form of corrosion, it is a result
of the unfavorable combination of three factors: An aqueous environment, generally considered
not very aggressive (pH between 5 and 10), a material deemed compatible with exposure
conditions and microorganisms whose presence is most often unexpected (Béranger &
Mazille, 2002). Since biocorrosion is not a new form of corrosion, the chemical reactions of
biocorrosion are the same as those of classical corrosion, however rates of corrosion are
increased.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 28

CHARACTERIZATION OF FACTORS CONTRIBUTING TO ENVIRONMENTAL


AGGRESSIVENESS

At Hydro-Québec, rock bolts are used in a variety of environments, such as:


1. underground : Wet wall, aggressive water or nonaggressive water, dry wall covered with
shotcrete…;
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2. at the surface : dry wall, submerged wall, wet wall, aggressive water, dust deposits,
nonaggressive water...
The appropriate level of protection must be determined by the environment’s aggressiveness
and the installation’s planned service life. A complete picture of the environment’s level of
aggressiveness on the support system according to parameters associated with corrosion
categories must be drawn. Factors contributing to rock bolt corrosion are shown in Table 2.

Table 2. Factors contributing to rock bolt corrosion (Source: ISO 9223)

Factor influencing corrosion Part of anchor protected by grout or Unprotected part of anchor (rod,
rate and type other coating (epoxy, etc.) head, nut, plate, etc.)
Humidity If moisture from the air or seepage Always in contact with air humidity,
water diffuses into the cement and therefore susceptible to corrosion
comes in contact with the bolt, it
promotes corrosion
pH of the water (pH) Dissolution of free lime and pH where corrosion is promoted
hydrolysis of silicates and depends on bolt material. For steel,
aluminates contained in the cement corrosion at acidic, neutral and
grout (e.g., pH 9) slightly alkaline pH
o
Water temperature ( C) Increased corrosion rate at higher temperature. Decrease in corrosion rate
as oxygen saturation decreases with increasing water temperature
Water flow rate (m3/s) Increase in flow rate brings aggressive ions and accelerates corrosion rate
Chlorides (mg/L) Increased risk of pitting corrosion
Sulfates (mg/L) Combine with tricalcium aluminate Increased risk of pitting corrosion
in cement to form Candlot’s salt
(ettringite), which disorganizes the
mortar by swelling it
Water containing dissolved Carbonation of grout cement and Not Applicable
carbon dioxide (mg/L) depassivation of steel
Oxygen (mg/L) If grout is permeable to oxygen, it If water flows over the bolt head:
promotes corrosion. oxygen supply promotes corrosion.
Water conductivity (µS/m or - Increase in corrosion rate with water
Ώ·m) or Total dissolves conductivity or increase in TDS
solids (TDS in mg/L)
Low water hardness Attack on grout by dissolution of Corrosive water because it does not
(dissolved Ca and Mg, total free lime and hydrolysis of silicates promote formation of a calcium
alkalinity (AT), carbonates, and aluminates carbonate passivating film
bicarbonates, CO2, OH- in
mg/L)
Presence of bacteria - Acceleration of aqueous corrosion
Cementitious grout The grout must not contain any Not Applicable
composition corrosive elements and must be
fairly impermeable to water and air.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 29

ENVIRONMENTAL AGGRESSIVENESS CLASSIFICATION

Atmospheric corrosivity rating

Environmental classification establishes the appropriate level of protection based on aqueous


and atmospheric environments’ aggressiveness, geochemistry of the site, bacterial environment,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

and the facility’s expected lifetime. ISO 9223 standard is widely used for classification of
atmospheric corrosivity. This standard presents two approaches for corrosivity determination
based on annual corrosion losses measured on reference metal specimens and estimation of
corrosivity based on the surrounding environment. In this second approach, the corrosivity
estimate can be normative, i.e., based on corrosion losses calculated on the steel through the
application of a “dose-response” function, or informative, i.e., the comparison between exposure
condition and description of typical atmospheric environments. ISO 9223 standard classifies
atmospheres into six corrosivity categories which applies to atmospheres encountered outside
buildings. Indoors, corrosivity is much lower because these areas are not subject to wind, rain, or
sudden changes in humidity or temperature. In underground powerplants, corrosivity classes can
vary greatly depending on bolt locations (e.g., surge chamber, engine room gallery).

Water corrosivity rating

Geotechnical applications such as road, bridge, tunnel and underground construction, and
mining, various water aggressiveness classifications include important parameters such as pH,
dissolved oxygen, calcium hardness, total alkalinity (AT) associated with calcium carbonate
precipitation, total dissolved solids (TDS), chloride ions Cl-, sulfate ions SO42-, rock quality,
temperature and water flow rate. Water and soil (or rock) corrosivity for anchor rods and
reinforcing rock bolts is addressed by the following committees: USA Transport Research Board
- National Research Council (Withiam et al., 2002), Design Manual for Roads and Bridges (The
Highways Agency (1999), the French Committee on Soil Mechanics and Foundation Works
(Habib, P., & Logeais, 1995) and the Post-Tensioning Institute (Post-Tensionning Institute,
2014), as well as authors Li & Lindblad (1999) and Roy (2016). The proposed aqueous
corrosivity classification has been adapted from these references and from research projects on
aging of reinforcing bolts to arrive at a proposed method for environments encountered in
hydroelectric facilities.
The main factors that influence water corrosiveness are pH values and dissolved oxygen
concentration. Figure 3 shows relationships between these ranges, the parameters to be
determined, and the associated failure mechanisms. In this project, most sites tested were mainly
low to moderate corrosivity, and sometimes high corrosivity. In the next section, corrosion
protection will be selected according to corrosivity levels.

CORROSION PROTECTION OF ROCK BOLTS

To render facilities and their support systems more durable, corrosion protection measures
must be adapted to both environmental corrosivity in which the rock bolts are located and
expected service life. Bolt corrosion protection measures are chosen on the basis aggressiveness
of two types of environments: atmospheric or aqueous. The bolt head, plate, ball joint and nut
can be protected against uniform and localized corrosion by protective paint. However, seepage

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 30

water can flow between its plate and the rock surface and potentially reach the rod, which then
becomes vulnerable to uniform and localized corrosion. Chloride concentration and sulfate ions,
as well as microbial corrosion, can increase environmental aggressiveness.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Decision Tree for Corrosion Protection of Reinforcing Bolts

Figure 4 shows the proposed decision tree with a choice between four levels of bolt corrosion
protection (no protection, Class 1, Class 2, Class 3, and Class 4 protection). The choice depends
on environmental aggressiveness, service life, and protection installation cost. In this document,
a reinforcing bolt is considered a permanent active bolt and must have corrosion protection if
service life exceeds 24 months.

Figure 4. Decision Tree for Corrosion Protection of Reinforcing Bolts (Source: ISO 9223)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 31

Proposed protection classes (1 to 4) are described in Table 3, for each zone of the bolt to be
protected, as defined previously (see figure 1). All these recommendations are made for
mechanically anchored rock bolts in our hydroelectric facilities. These bolts are usually
subjected to significant mechanical stress in tension, so that surface corrosion, even if localized
at one point, may cause the rod to crack suddenly, even if corrosion has not caused a significant
reduction in the cross-section of the rod. New facilities will have to consider environmental
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

aggressivity and associated classes before designing rock support systems.


This protection strategy is not necessarily adapted for the other components frequently used
in support systems, i.e., pins and wire mesh. Mesh is usually covered with a galvanizing layer
that temporarily protects them against corrosion. Pins are not galvanized and neither protected
with grout.

Table 3. Aqueous and Atmospheric Corrosion Protection Classes According to the Zones of
the Reinforcing Rock Bolt

Protection Rock bolt zones (see Figure 1)


classes
Zone 1[1] Zone 2 Zone 3 Zone 4
Class 1 Coating (paint) adapted to CT bolt with Cement grout adapted to CT bolt with
a highly aggressive protective sleeve the aggressiveness of the protective
environment environment over the sleeve
entire length of the bolt
Class 2 Coating (paint) adapted to Soft petrolatum Cement grout adapted to Galvanization
a moderately aggressive primer and mastic the aggressiveness of the
environment environment over the
entire length of the bolt
Class 3 Galvanization Soft petrolatum Cement grout adapted to Galvanization
primer and mastic the aggressiveness of the
environment over the
entire length of the bolt
Class 4 No protection Soft petrolatum Cement grout adapted to Black steel rod
primer and mastic the aggressiveness of the
environment over the
entire length of the bolt
[1] For existing bolts, apply protections for this area only.

CORROSION PROTECTION: PAINTS

This section deals solely with protection methods (plate, washer, and nut) with a flexible
coating to prevent the metal from coming into contact with moist air or corrosive seepage water
(Protection Classes 1 and 2). Hydrophobic products, such as pastes, waxes, greases and paints,
can be used and applied after bolt installation. Organic coatings, or organic paints, are most
widely used. Adequate and lasting corrosion protection depends on the following factors: paint
systems, design of the structure, steel condition before preparation, degree of surface preparation,
quality of surface preparation work, condition of assemblies, edges and welds before preparation,
application work, conditions under which the application is carried out and post-application
exposure conditions. For more information, please refer to ISO 9223 and ISO 12944.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 32

Recommended coatings for highly aggressive environments (Class 1)

Bolts in a highly aggressive environment may be exposed for decades to acidic seepage water
(pH < 5) and concentrated sulfate ions (> 750 ppm). If seepage water dries out, pH and sulfate
content can increase considerably through a concentration effect. Note: take these parameters
into account when choosing the type of coating to use. ISO 12944-6 (BS-EN-ISO-12944-6,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1998) was used in the laboratory to qualify the resistance of recommended coatings to sulfuric
acid, using immersion procedure from ISO 2812-1 (BS-EN-ISO-2812-1, 2007). Figure 5 shows
authors’ recommendations for coating application to bolt heads in highly acidic environments.

Figure 5. Recommendations for Application of Coatings to Bolt Heads in Highly Aggressive


Environments

Due to prolonged exposure of bolts to seepage water in the plant, an immersion time of 648
hours was chosen for plates and 168 hours for mounts (mounts have a lower quality surface
preparation). For prolonged exposition, epoxy novolac coatings are more resistant than an epoxy
polyamine and novolac mixture. However, application techniques and tolerance to application
constraints for novolac coatings are more restrictive. The recommended coatings in this section
are still being evaluated in real-world environments in two underground powerplants.

Recommended coatings for moderately aggressive environments (Class 2)

Figure 6 shows authors’ recommended products for painting rock bolt heads in moderately
aggressive environments. This selection was based on recommendations from various specialists,
providing these criteria for exposure conditions of the bolts: moderate atmospheric corrosion
with the presence of seepage water in a few locations.

Figure 6. List of Recommended Coatings for Moderately Aggressive Environments

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 33

CONCLUSION

Function and serviceability of support systems in existing facilities must be ensured.


Investigation of the extent of corrosion, its effect on bolt serviceability and possible causes of
premature rock bolts aging in a hydroelectric setting were central concerns in this industrial
project. Results of this project showed that corrosivity levels in our tested facilities varied from
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

low to moderate, with a few sites exhibiting high levels. Protection classes needed in our
facilities ranged from 1 to 4, with classes 1 and 2, coating types and systems were tested and
recommended.
The results of this industrial project allowed Hydro-Québec to have a better understanding of
the impact of rockbolt corrosion on support systems in their numerous facilities. Moreover, the
technical guide is a reference document for design, construction, maintenance and monitoring of
these systems. Several decision trees and practical solutions allow to ensure the safe operation of
support systems in our facilities.

ACKNOWLEDGEMENT

The authors wish to thank Hydro-Québec Management for subsidizing this industrial
research project and for authorizing this publication. We personally wish to thank Kaveh Saleh, a
retired Hydro-Québec engineer who led this project.

REFERENCES

Béranger, G., and Mazille, H. (2002). Corrosion des métaux et alliages : mécanismes et
phénomènes. Paris: Hermes Science Publications.
Bertolini, L. (2004). Corrosion of steel in concrete: prevention, diagnosis, repair. Weinheim,
Wiley-VCH.
Cigna, R., Andrade, C., Nürnbergrt, U., Polder, R., Weydert, R., and Seitz, E. (2003). Corrosion
of Steel in Reinforced Concrete Structures – Final Report (E. C. i. t. F. o. S. a. T. R. (COST),
Trans.) COST Action 521 (pp. 238). Bruxelles: European Commission.
Crosky, A., Fabjanczyk, M., Gray, P., Hebblewhite, B., and Smith, B. (2002). Premature Rock
Bolt Failure (A. C. A. R. Program, Trans.): The University of New South Wales Mining
Research Centre, SCT Operations Pty Ltd, Ground Support Services Pty Ltd.
Dorion, J. F. (2013). La Corrosion du Soutènement Minier. (Thèse de Doctorat. Université
Laval. Québec, Canada).
Gamboa, E., and Atrens, A. (2003). Environmental influence on the stress corrosion cracking of
rock bolts. Engineering Failure Analysis, 10(5), 521-558.
Habib, P., and Logeais, L. (1995). Tirants d’ancrage : Recommandations concernant la
conception, le calcul, l’exécution et le contrôle. Recommandations T.A. 95. In Comité
Français de la Mécanique des Sols et des Travaux de Fondations (Ed.), (pp. 184): Eyrolles.
Landolt, D. (1993). Corrosion et chimie de surfaces des métaux. Lausanne, Suisse: Presses
polytechniques et universitaires romandes.
Li, C., and Lindblad, K. (1999). Corrosivity classification of the underground environment. Dans
Rock support and reinforcement practice in mining: proceedings of the international
symposium on ground support. Kalgoorlie, Western Australia, pp 15-17. Rotterdam, Pays-
Bas: Balkema Pub., AA/Taylor & Francis.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 34

Pourbaix, M., and de Zoubov, N. (1974). Section 12.1: Iron. In M. Pourbaix (Ed.), Atlas of
Electrochemical Equilibria in Aqueous Solutions (pp. 307-321). Houston, Texas USA: Natl.
Ass. of Corrosion Engineers.
Post-Tensionninig Institute. (2014). Recommendations for Prestressed Rock and Soil Anchors.
Norme PTI DC35.1-14. Londres, Royaume-Uni.
Revie, R. W. (2011). Uhlig’s Corrosion Handbook (3rd Edition). New York, USA : John Wiley
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

& Sons.
Roy, J. M., Preston, R., and Bewick, R. P. (2016). Classification of Aqueous Corrosion in
Underground Mines. Dans Rock Mechanics and Rock Engineering, Feb., 2016.
The Highways Agency. (1999). Design Manual for Roads and Bridges (DMRB) Volume 2,
section 1, part 7 – Use of rockbolts. Écosse, BA -80/99.
Withiam, J. L., Fishmann, K. L., Gaus, M. P., and T. R. B. N. C. H. R. Program. (2002).
Recommended practice for evaluation of metal-tensioned systems in geotechnical
applications. Washington, D.C., National Academy Press.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 35

Use of Standard Penetration Test (SPT) to Determine Raveling Index

Timothy Copeland1; Ryan Shamet2; Jinwoo An3; Kyungwon Park4; and Boo Hyun Nam5
1
Dept. of Civil, Environmental, and Construction Engineering, Univ. of Central Florida, Orlando,
FL
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
School of Engineering, Univ. of North Florida, Jacksonville, FL
3
Univ. of Mount Union
4
Dept. of Civil Engineering, Kyung Hee Univ., Giheung-gu, Yongin, Gyeonggi-do, Republic of
Korea
5
Dept. of Civil Engineering, Kyung Hee Univ., Giheung-gu, Yongin, Gyeonggi-do, Republic of
Korea (corresponding author). Email: boohyun.nam@khu.ac.kr

ABSTRACT

In the East-Central Florida region, where there is relatively deeper depth of bedrock than
north/west of Florida, cone penetrometer tests (CPTs) have been popularly used for sinkhole
geotechnical investigation. From the CPTs, an index known as the raveling index (RI) was
developed. This raveling index is used to better understand the status of raveling, thus,
quantitatively evaluating the susceptibility of sinkholes. However, standard penetration test
(SPT) is still the most common subsurface exploration method in geotechnical engineering
practice. In areas of shallow bedrock/limestone or where CPT testing is unavailable, SPT testing
is necessary, often the case in West-Central Florida. In this paper, a method for calculating the
raveling index in known sinkhole locations using SPTs was developed. An analysis of the RIs
calculated using CPTs was performed to compare with those calculated using SPTs. The
outcome of this analysis is a preliminary result showing a potential benefit of SPT in sinkhole
investigation and also a limited correlation between the RI calculations. Subsequently, there may
be a possibility of using SPTs in tandem with CPTs for future projects to be effectively used to
investigated for sinkhole risk. This study was applied to various sinkhole locations throughout
Central and northern Florida.

INTRODUCTION

When groundwater infiltration rapidly recharges the surficial aquifer after a rush of rainfall,
such as during a storm, the restricted aquifer (i.e., the Floridan aquifer) recharges more slowly
due to the occurrence relatively low permeable soils between them. This case of
hydrogeomorphology results in a dramatic increase in vertical head difference in the
groundwater regime (Xiao et al. 2016; Shamet et al. 2020). Internal soil erosion occurs under the
ground surface as a result of the head differential. A sinkhole can form when erosion becomes
severe enough and the resulting eroded void is large enough that it can no longer support the
overlying soil, also known as the raveled zone (Nam and Shamet 2020). Sinkhole creation (cover
collapse type) is a fast-moving process that can cause significant personal and property harm.
Physical model tests (Perez et al. 2016), sinkhole raveling chart and vulnerability index (Nam
and Shamet 2020, Shamet et al. 2018), stability analysis via numerical simulation (Soliman et al.
2019), identification and analysis of sinkhole factors (Nam et al. 2020), remote sensing (Kim et
al. 2019), and piezometric studies (Copeland et al. 2021; Copeland et al 2022) have all been used
by researchers in Florida to quantify sinkhole vulnerability (Kim et al. 2020).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 36

One of the physical test metrics used to understand sinkhole development is the Raveling
Index, RI (Foshee and Bixler 1994). While this is a great metric, it does have limitations. One of
these limitation is that it is dependent on the profile developed from a CPT. While CPTs are
becoming more common, they are not always the chosen method of soil investigation. Typically,
this is due to one of two factors the depth of bedrock and the availability of equipment. Where
the bedrock is shallow a CPT test is less likely to occur because it will be unable to penetrate the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

bedrock layer and there the soil profile would be incomplete. In this case other methods of soil
investigation are used, most often a standard penetration test, SPT. The SPT is a soil
investigation technique that determines the strength of soil at a specific depth based on the
penetrative resistance of the soil. The SPT has been used to determine sinkhole activity both
independently by Zisman 2019 and in tandem with GPR by Kiflu et al. 2013. Understanding this,
the authors made an attempt to calculate the raveling index based on the SPT profile and
compare the results with that of nearby CPT profiles.

STUDY AREAS

This study included multiple sites throughout the state of Florida. Each of the sites were
investigated by the Florida Department of Transportation throughout North and Central Florida.
Ground investigation data from eleven total sites were used in this study, Figure 1. From the
eleven sites a total of 26 comparisons between SPTs and CPTs were evaluated. A list of the
number of these comparisons can be found in Table 1. The number of comparisons was
determined based on the number of known comparable quantities, where the SPT & CPT were
either located in close proximity or within the relative location of the sinkhole. Using this criteria
and reasonable engineering judgement, each comparison was then made and used to evaluate the
effective of the SPTs for use in calculating the RI.

Figure 1. Site Location Map

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 37

Table 1. Study Locations

Number of
Site Info/File Name SPT/CPT
Comparisons
I-4 CR 46 Depression 2002 1
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

I-4 Maitland Blvd Depression 2004 5


I-4 Rest Area Polk County 2011 1
I-75 Northbound Depression Alachua 2005 1
PSI Pasco County Land O' Lakes 2018 5
SR 44 Depression 2014 1
SR 535 Meadow Creek 2006 1
SR 55 Taylor County 2021 6
US 19 Sealawn 2011 2
US 27 Polk County 2010 North 2
US 27 Polk County 2010 South 1
Total 26

DATA ANALYSIS

First proposed by Foshee and Bixler (1994) the raveling index, RI, is a technique used to
assess the sinkhole danger based on a known cone penetration test, CPT, profile, Figure 2. The
index is calculated by dividing the depth to the top of the raveled zone, tover, by the thickness of
the raveled soil layers, travel.
𝑡𝑟𝑎𝑣𝑒𝑙
𝑅𝐼 = (1)
𝑡𝑜𝑣𝑒𝑟

The RI indicates how much erosion has happened in the overburden sandy soils. The larger
the RI, the more erosion indicated in the soil profile.
The start and end of the raveling zone for the CPTs are typically determined when a tip
resistance is particularly low as identified in Figure 2. Similarly, the raveling zone for the SPT
can be determined when the N-value or blow count is significantly low, typically a blow count
below 5 to weight of hammer. The start and end points of the raveling zone are easier to identify
using the CPT because unlike the SPT it is a continuous test. Therefore, the first step of the
analysis is to determine the raveling zone of each CPT. Once determine, the raveling zones of the
SPTs can be determined. Using the extent of each raveling zone, the RI for each test can be
calculated and the appropriate comparison between the two tests can be made.

SPT & CPT Comparison

As previously mentioned, the CPT is a more prominent testing method for geotechnical
investigation in North and Central Florida. Given this, the study sites generally consisted of
multiple CPTs and fewer SPTs. Figure 3 shows one of the sites used to compare RI. The I-4
Maitland Blvd depression consisted of 12 total CPTs and 5 SPTs. One comparison performed
was between CPT-6 and SPT boring B5. This particular case identified two borings located in
almost the same location and provided a good profile comparison. The raveling zone identified

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 38

from the CPT, 57 feet to 111 feet, and the SPT, 60 feet to 130 feet, were comparable with a
similar RI calculation of 0.95 and 1.17 respectively. While it may be noted that there is a large
spike at a depth of 90' for the SPT, this was considered to be an anomaly due to a possible
drilling induced error or other interference. Figure 4 shows a different example of a SPT/CPT
comparison and how it was an example of an outlier. SPT B-1 and CPT-6 while located
approximately 9 feet apart, are in completely different zones as related to the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

sinkhole/depression. The SPT is located inside the depression while the CPT is located outside.
This resulting difference has an extreme effect on the raveling index, with the CPT RI of 0.33
and SPT RI of 13.0. Therefore, this comparison was not considered for the analysis and
additional criteria such as relative location to the sinkhole was also considered.

Figure 2. Example SPT and CPT Profile with identified raveled and non-raveled zones

Figure 3. I-4 Maitland Blvd Depression (Note: C-#: CPT boring, B-#: SPT boring)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 39
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. I-75 NB Depression Alachua County (Outlier – CPT and SPT not in ‘similar’
location could not compare results)

RESULTS & DISCUSION

Table 2 shows the RI calculation results for both the SPTs and CPTs. Each row of the table
indicated the location of the test was well as the identified SPT and CPT that were compared. It
should be noted that there are 26 unique SPTs, however, due to the location of these tests, the
same CPT may have been used as a comparison for multiple SPTs.
The results of the comparison, CPT RI vs SPT RI, are graphed, Figure 5. The initial
evaluation shows a linear trend that favors the CPT RI calculation, indicating that the CPT
calculation may provide a higher RI value than that from the SPT. A larger sample would be
necessary however to prove that this trend is substantial.

Figure 5. SPT RI vs. CPT RI

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 40

Given that these investigative tests took place in and around the sinkholes/depression a
second approach was considered based on the tests’ location. Figure 6 shows the comparison
based on these locations; IN - those located within the specified sinkhole/depression, EDGE-
those located near the extent of the sinkhole/depression (typically within 2 m of specified
boundary), and OUT – those located outside the sinkhole/depression. The test located IN showed
a more significant linear trend than those on the EDGE or OUT. Typically, these test locations
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

are the closest in proximity and therefore it would make sense that these tests would have the
best correlation.

Table 2. SPT & CPT Data Collection

Closest
Site Info/File Name SPT ID SPT RI CPT RI
CPT ID
I-4 CR 46 Depression 2002 SPT-2 0.35 CPT-1 0.58
I-4 Maitland Blvd Depression 2004 SPT-1 1.75 CPT-3 2.87
I-4 Maitland Blvd Depression 2004 SPT-2 4.00 CPT-5 10.00
I-4 Maitland Blvd Depression 2004 SPT-3 4.33 CPT-5 10.00
I-4 Maitland Blvd Depression 2004 SPT-4 0.25 CPT-8 0.18
I-4 Maitland Blvd Depression 2004 SPT-5 1.17 CPT-6 0.95
I-4 Rest Area Polk County 2011 SPT-5 0.77 CPT-3 0.60
I-75 Northbound Depression Alachua 2005 SPT-3 0.50 CPT-4 0.26
PSI Pasco County Land O' Lakes 2018 SPT-10 0.13 CPT-5 0.14
PSI Pasco County Land O' Lakes 2018 SPT-4 0.16 CPT-3 0.17
PSI Pasco County Land O' Lakes 2018 SPT-5 0.21 CPT-3 0.17
PSI Pasco County Land O' Lakes 2018 SPT-8 0.11 CPT-5 0.14
PSI Pasco County Land O' Lakes 2018 SPT-9 0.20 CPT-5 0.14
SR 44 Depression 2014 SPT-1 0.42 CPT-1 0.51
SR 535 Meadow Creek 2006 SPT-5 0.33 CPT-5 2.37
SR 55 Taylor County 2021 SPT-1 0.31 CPT-1 1.86
SR 55 Taylor County 2021 SPT-2 1.82 CPT-3 2.17
SR 55 Taylor County 2021 SPT-3 4.00 CPT-2 2.00
SR 55 Taylor County 2021 SPT-4 1.27 CPT-2 2.00
SR 55 Taylor County 2021 SPT-5 0.13 CPT-2 2.00
SR 55 Taylor County 2021 SPT-7 2.14 CPT-1 1.86
US 19 Sealawn 2011 SPT-1 0.83 CPT-4 0.56
US 19 Sealawn 2011 SPT-2 0.36 CPT-1 0.87
US 27 Polk County 2010 North SPT-B1 0.53 CPT-6 0.62
US 27 Polk County 2010 North SPT-B2 0.33 CPT-6 0.62
US 27 Polk County 2010 South SPT-SK1 0.16 CPT-1 0.46

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 41
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 6. SPT RI vs CPT RI based on test location (IN, EDGE, and OUT)

In addition to addressing the tests’ location an analysis based on the RI calculation was
completed. This analysis is significant because understanding what the RI means is important. If
the raveling index is low (less than 1.5) then there is likely little to nothing to done while if the
raveling index is high (greater than 4) then some type of mitigation effort will need to be
performed. It then becomes difficult to understand the mid-range (1.5 to 4) and making sure that
the RI calculations from both tests are consistent. Figure 7 shows that this is not the case. While
the mid-range RI calculations from the CPTs are consistent the SPT calculations range from 0.1
to 2.15 indicating a larger difference in possible necessity for remediation. Therefore, using the
SPT to calculate the RI for those in the mid-range may be too difficult based on the limitation of
the tests located in the section below.

Figure 7. SPT RI vs CPT RI based on RI calculation (Low, Mid, and High)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 42

Limitations

It should be noted that there are multiple limitations. This method was somewhat subjective
when determining the boundaries for the RI calculation of the SPTs by comparing them to the
boundaries of the CPTs raveling zone. This influence was limited as much as possible using a
‘blind’ process to determine the initial raveling zones of the SPTs. However, the raveling zone
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

for the SPTs is more difficult to determine because, unlike the CPT data, the SPT data is
segmented every 2.5 to 5.0 feet in depth. Therefore, precisely determining the start and end of
the raveling zones from the SPT data can become a problem. This could cause some level of
variation in determining the actual raveling zone when compared to that of the CPT. More data
collection and statistical analysis is necessary.

CONCLUSIONS

This paper presents a preliminary analysis of the comparison of SPT and CPT in sinkhole
vulnerability assessment. The Raveling Index (RI) calculation, based on an established CPT
method, was employed to quantitatively compare the accuracy of the proposed SPT method. The
preliminary outcome from this paper shows a potential but limited correlation between the RI
computations. The comparably lengthy sample interval appears to limit the SPT's usefulness for
estimating RI, even if it may be accessible on a drill site when bedrock and coring or other
investigation is required. As a result, it may be possible to employ SPTs in conjunction with
CPTs for future projects to properly assess sinkhole risk. However, a more comprehensive
research approach using a larger data set is recommended for independent use of SPTs for
sinkhole assessment.

ACKNOWLEDGEMENT

The authors would like to thank Dr. David Horhota from Florida Department of
Transportation (FDOT) State Materials Office (SMO) for his guidance and support.

REFERENCES

Copeland, T. R., Nam, B. H., Kim, Y. J., and Han, H. S. (2021). Use of X-Bar and R Control
Chart Methods on Long-Term Piezometer Data for Sinkhole Assessment. Geo-Extreme
Conference, 127-136.
Copeland, T. R., Nam, B. H., Shamet, R., Youn, H., and Jung, Y. H. (2022). Investigation on the
Effects of a Hurricane on the Piezometric Head and Lateral Hydraulic Gradient of Network
of Piezometric Sensors in Florida’s Sinkhole Study. Geo-Congress Conference, 495-504.
Foshee, J., and Bixler, B. (1994). Cover Subsidence Sinkhole Evaluation of State Road 434,
Longwood, Florida. 120(11). pp. 2026–2040.
Kiflu, H., Kruse, S., and Wightman, M. (2013). Statistical analysis of GPR and SPT methods for
sinkhole investigation in covered karst terrain, West-Central Florida, USA. In 13th Sinkhole
Conference (pp. 247-253). National Cave and Karst Research Institute.
Kim, Y., Nam, B. H., Shamet, R., Soliman, M., and Youn, H. Development of Sinkhole
Susceptibility Map of East Central Florida. Natural Hazards Review 21 (4), 04020035.
Kim, Y., Nam, B., and Youn, H. Sinkhole Detection and Characterization Using LiDAR-Derived
DEM with Logistic Regression, Remote Sensing 11 (13), 1592.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 43

Nam, B. H., Kim, Y., and Youn, H. Identification and quantitative analysis of sinkhole
contributing factors in Florida’s Karst, Engineering Geology 271 (2020) 105610.
Nam, B. H., and Shamet, R. (2020). A preliminary sinkhole raveling chart. Engineering Geology
268 (2020) 105513.
Perez, A., Nam, B., Alrowaimi, M., Chopra, M., Lee, S., and Youn, H. (2016). Experimental
Study on Sinkholes: Soil–Groundwater Behaviors under Varied Hydrogeological Conditions.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

J. Test. Eval. 45, 208–219.


Shamet, R., Soliman, M., Kim, Y., Copeland, T., and Nam, B. H. (2020). Sinkhole Investigations
After Hurricane Irma. In NCKRI Symposium XX 15th Sinkhole Conference (pp. 1-10).
Shamet, R., Horhota, D., and Nam, B. (2018). “Development of a point-based index for Sinkhole
Vulnerability Evaluation in Central Florida’s Karst Terrain.” International Foundation
Construction Equipment Expo, ASCE, Orlando, FL, 212-221.
Soliman, M. H., Shamet, R., Kim, Y. J., Youn, H., and Nam, B. H. (2019). Numerical
investigation on the mechanical behavior of karst sinkholes. Environmental Geotechnics 0,
1–15.
Xiao, H., Kim, Y. J., Nam, B. H., and Wang, D. (2016). Investigation of the impacts of local-
scale hydrogeologic conditions on sinkhole occurrence in East-Central Florida, USA.
Environ. Earth Sci. 75, 1274.c.
Zisman, E. D. (2019). Determination of Sinkhole Activity in Granular Sediments in West-
Central Florida. Journal of Testing and Evaluation, 47(4), 2512-2530.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 44

Integration of Downhole Data Reduction Techniques for Determination of 𝑽𝑺 Profiles

Ayush Kumar1 and Anbazhagan Panjamani, Ph.D.2


1
Research Scholar, Geotechnical Engineering Group, Dept. of Civil Engineering, Indian Institute
of Science, Bengaluru, Karnataka, India. Email: ayushkumar@iisc.ac.in
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Associate Professor, Geotechnical Engineering Group, Dept. of Civil Engineering, Indian
Institute of Science, Bengaluru, Karnataka, India. Email: anbazhagan@iisc.ac.in

ABSTRACT

The downhole method is one of the most widely used in situ seismic tests to determine shear
wave velocity (𝑉𝑆 ) profiles for site characterization and site response analysis. Although easy to
perform in the field, it has always been a tedious task to accurately interpret the downhole data
based on the arrival times of shear waves at the borehole receiver. Commonly used methods are
the direct, interval, and Snell’s Raypath methods. Snell’s Raypath method has been known to
provide the best results for downhole processing at multilayered subsurface sites as it considers
the actual travel path of waves considering the refraction along the wave travel path. The direct
method is considered better when the soil layers are uniform with reduced chances of refraction.
In this study, 𝑉𝑆 profiles are estimated for two boreholes 100 m deep at a deep silty deposit
underlain by sandstone rocks. The Snell’s Raypath method was observed to be more suitable for
shallow subsurface to capture a greater variation in properties and need for a higher resolution in
shallow subsurface. The direct method was found suitable for deeper layers where dense soil and
rock layers are located, which show more uniform layering with depth.

INTRODUCTION

Shear wave velocity (𝑉𝑆 ) is the key parameter in the site response analysis or seismic site
characterization. It is one of the most useful engineering properties of soil and rock because of its
direct relationship with the shear modulus of the material. Thus, 𝑉𝑆 profiling is an important
parameter for major engineering studies. These profiles are measured in the field using seismic
methods. The depth of investigation generally varies between 30 to 150 m. However, deeper
profiling up to 500 – 600 m has also been carried out recently (Stokoe et al. 2017, Hwang et al.
2018). There have been several methods developed to interpret the downhole data and estimate
accurate 𝑉𝑆 profiles. However, there is uncertainty over the applicability of these methods
because these all methods determine 𝑉𝑆 profiles in different ways. There is no consensus on
using these methods in an integrated manner to obtain reliable profiles as each method has its
own benefits and shortcomings. The widely followed test standard D7400 – 2019 suggests the
use of direct and interval methods, which are applicable when refraction in subsurface can be
neglected or is expected to be minimum. Thus, in this paper, an attempt has been made to
understand differences in profile obtained from two commonly used interpretation methods and
combine them together. The major objective is to highlight the importance of an integrated
approach in downhole data interpretation. In this study, the downhole method commonly used to
measure 𝑉𝑆 is described. Methods used for interpretation of arrival time of generated waves and
processing of those arrival times for calculation of 𝑉𝑆 are discussed. Finally, downhole seismic
tests were carried out at two locations in the East Indian Coastal region and compared with
borehole profile and MASW results performed at the same location.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 45

DOWNHOLE TEST

The downhole test is widely used for the in-situ determination of shear (S-) and
compressional (P-) wave velocities of soil and rocks. The downhole method estimates the
velocity of body waves in different subsurface layers by measuring the arrival time of waves
from the source on the surface to the receiver/s at different depths in one borehole. The
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

downhole test is economical because of a single borehole, easy procedure, and use of surface
source as simple as a sledgehammer and a wooden plank. Another added advantage is that
samples are obtained during drilling, which can later be used for a detailed study of soil and rock
properties.

Figure 1. Schematic of a typical downhole test showing surface source and two receiver
positions in the borehole

The schematic of the test is shown in Figure 1. The test consists of an impulsive or vibratory
source for the generation of P- and S-waves on the surface. A static load is placed over the
source to ensure firm contact between the source and the ground. A 3- (or more) channel receiver
is placed in the borehole with a clamping mechanism to have firm contact with the borehole
lining. After each acquisition, the receiver is progressed to the next recording depth. The
acquisition interval can be 1 m or 1.5 m or can be decided based on the total depth of the
borehole as well. For detection of the arrival of S-waves, polarity reversal of S-waves (when the
sledgehammer hits the shear beam on the opposite sides, or the vibration source is excited in two
opposite directions) is employed. This is discussed in the next section.
Data Processing. For the determination of 𝑉𝑆 profiles, arrival times of body waves should be
determined first. To determine the arrival time of an S-wave at a specific depth, reverse polarised
S-waves are used, as shown in Figure 2. This method is termed as crossover method. It is a
general observation that S-waves show a change in polarity when the excitation direction of the
source is reversed. The section of wave signal before the arrival of the S-wave is mostly P-wave,
which does not show polarity reversal. So, S-waves can be clearly identified. Hence, this
technique is utilized. After the arrival time of the wave at all the depths is obtained, data
reduction techniques are applied, as discussed below.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 46
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Detection of S-wave arrival time (t1) by Crossover method

Direct Method. The direct method considers a straight travel path from the source to the
receiver. This method can be useful when the source distance is sufficiently small compared
to the depth of acquisition and refraction because the multi-layer soil profile can be
neglected. For processing, first, the measured travel time in the inclined ray path is corrected
to the vertical travel time as in Eq. 1 (Mok 1987, Batsila 1995, Kim et al. 2004).
𝑡
𝑡𝑐 = 𝐷 𝑅 (1)

where, 𝑡𝑐 = Vertical (or Corrected for Vertical) arrival time, D = depth of acquisition, R = source-
receiver distance.
After plotting 𝑡𝑐 vs. D, the velocity of each layer can be found by measuring the slope of
straight lines which best fit these points in their respective layers.
Interval Method. The Interval method uses the difference in travel time of waves
between two depths of acquisition. The wave velocity is given by Eq. 2 (Batsila 1995; Mok
1987, Kim et al. 2001).
𝑅2 −𝑅1
𝑉= (2)
𝑡2 −𝑡1

where, 𝑅1 = source-receiver distance for depth of acquisition 𝑅2 = source-receiver distance for


lower depth of acquisition, 𝑡1 = travel time for upper depth, 𝑡2 = travel time for lower depth.
The Interval method is simple to use. However, it does not consider the velocities of all the
layers along the ray path and fails when 𝑡2 < 𝑡1, i.e., travel time to the lower depth is lower than
the travel time to the upper depth. This can occur in profiles with high velocity contrast.
Snell's Raypath Method. The previous two methods considered a straight line, inclined
travel path between source and receiver. To overcome this limitation, computations
considering Snell's Law along the travel path were developed, resulting in Snell's Refracted
Raypath method (Joh and Mok 1998, Bang 2001). This method assumes a refracted ray path
based on Snell's Law between the subsurface layers (Figure 3). For the first layer, 𝑉1 can be
calculated directly based on straight raypath length and arrival time as the uniform layer is
assumed. Based on the travel time and source-receiver geometry for the subsequent layers,
the following equations are solved to calculate velocity (Kim et al. 2004).

sin 𝜃𝑖,1 sin 𝜃𝑖,2 sin 𝜃𝑖,𝑗 sin 𝜃𝑖,𝑖


Snell's Law: = =⋯= = (3)
𝑉1 𝑉2 𝑉𝑗 𝑉𝑖

𝑍1 tan 𝜃𝑖,1 + 𝑍2 tan 𝜃𝑖,2 + ⋯ + 𝑍𝑗 tan 𝜃𝑖,𝑗 + ⋯ + 𝑍𝑖 tan 𝜃𝑖,𝑖 = 𝑆 (4)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 47

𝑍
𝐿𝑖,𝑗 = 𝑐𝑜𝑠 𝑗𝜃 (5)
𝑖,𝑗

Then, velocity can be calculated as


𝑍𝑖
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

𝐿𝑖,𝑖 𝑐𝑜𝑠 𝜃𝑖,𝑖


𝑉𝑖 = = 𝐿𝑖,𝑗 (6)
𝑡𝑖,𝑖 𝑡𝑖 −∑𝑖−1
1 𝑉
𝑗

where, 𝐿𝑖,𝑗 = Length of travel path for ith receiver in jth layer, 𝑡𝑖,𝑗 = Time of travel along ray path
for the ith receiver in the jth layer, 𝜃𝑖,𝑗 = Angle with normal for ith receiver's ray path in jth layer

Figure 3. Illustration of Snell's Raypath Method for a three-layer model (after Kim et al.
2004)

To start with these equations, First an initial value of 𝑉𝑖 needs to be assumed, which can be
𝑅
the average velocity given by 𝑉𝑖 = 𝑖⁄𝑡 (𝑅𝑖 is the straight line distance between source and
𝑖
receiver at concerned depth, 𝑡𝑖 is the arrival time at concerned depth). These equations are solved
by iteration. 𝑉𝑖 is updated after every iteration. The iteration is continued till the difference
between the assumed and the calculated velocity at the step reduces below the defined lower
limit (e.g., 0.01%) (Kim et al., 2004). In the present study, the calculations were performed using
MATLAB 2021a software package.
It has been observed that these different methods result in different 𝑉𝑆 profiles which often do
not agree with subsurface layering. In order to understand the differences obtained due to
application of these different methods, and how to integrate them to utilize the advantages
provided, an integrated approach is proposed in this study.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 48

FIELD STUDY AND DATA ACQUISITION

This study is carried out near the eastern coast of India over a silty deposit underlain by
sedimentary rock layers. Two boreholes up to a depth of 100 m were drilled using hydraulic
rotary drilling equipment accompanied by the Standard Penetration Test (SPT) for sample
collection and general characterization of the subsurface. The boreholes were then prepared
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

for the downhole test as per ASTM 7400 – 2019. The downhole test setup included a wooden
shear beam with steel caps at the ends, a sledgehammer to generate an impulse, and a metal
plate for vertical impulses. The beam was kept at a distance of 2.5 m from the borehole. The
impulses were given in opposite directions to reverse the S-wave's polarity to easily identify
its arrival. Geode Seismograph (Geometrics) was used for data acquisition along with a
BGK7 borehole receiver (Geotomographie GmBH). Impulses were recorded at every 1 m
depth interval.

Figure 4. Typical S-wave signal traces at 10m depth intervals at site DH 2

MASW tests were carried out at both the test location using Geode seismograph and 24
nos. 2Hz geophones. A sledgehammer and a metal plate were used as source. Geophone
interval was kept as 1m and source was placed at 10m from the nearest geophone. The 𝑉𝑆
profile was determined considering a 10-layer subsurface model.

RESULTS

Before arrival time detection, a low pass filter of 100 Hz was applied to remove the noise
from the wave records. Then, both the direct method and Snell's Raypath method were used
for processing S-wave arrival times. A typical waterfall plot of the acquired waveforms at
10m intervals at the DH 2 site is shown in Figure 4. Vertical arrival times (𝑡𝑐 ) are plotted
against depth for both the boreholes in Figure 5(b) and Figure 6(b). Firstly, the direct method
is used to estimate the 𝑉𝑆 . The direct method was observed to miss a detailed profile for

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 49

shallow depth where variability in soil properties is known beforehand from borehole drilling
and soil sampling. Hence, Snell's Law Raypath is carried out for shallow depth to assess the soil
stiffness with a better resolution and later integrated with the 𝑉𝑆 profile from the direct method.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5. (a) Soil profile from borelog (b) Vertical arrival time for the downhole test at DH
1, (c) 𝑽𝑺 from direct method and Snell's Raypath method, (d) 𝑽𝑺 from Integrated method
and MASW along with SPT N-values

If Snell's Raypath 𝑉𝑆 profile is extended to deeper layers, the results can be misinterpreted
because of two reasons: (1) there are very low chances for refraction as the wave travel path
becomes essentially straight/vertical with an increase in depth (Kim et al., 2004), and (2) The
variation in stiffness reduces and layers tend to be more uniform, while Snell's Raypath method
can result into highly fluctuating 𝑉𝑆 profiles which is not possible in uniform rock/soil layer.
Hence, Snell's Raypath method in shallow depth and the Direct method in deeper layers might be
a better combination to use for more reliable 𝑉𝑆 profiles for deep boreholes.
The depth up to which Snell's Law can be considered useful depends upon the test geometry
and soil profile determined from the borelog. 𝑉𝑆 profiles from both the methods are shown in
Figure 5(c) and Figure 6(c), and combined in Figure 5(d) and Figure 6(d), along with 𝑉𝑆 profile
from MASW and subsurface profile from borelog (Figure 5(a) and Figure 6(a)). The 𝑉𝑆 profile
from the MASW test agrees with the downhole results in general, except for a few outlier layers
where the difference is high.
N-values obtained from SPT tests at the site are also presented in Figure 5(d) and 6(d). It can
be observed that in general, N-values follow the trend similar to the 𝑉𝑆 profile, except a few
outliers. It is to be noted that hammer energy was not carried out during SPT, which majorly
influences the N-values (Anbazhagan et al. 2021). Hence corrected N-values could not be
determined. There are a few layers where lower N-values correspond to the higher 𝑉𝑆 when
compared to subsequent deeper or shallower layers, and vice versa. However, it might not be
necessary that an increase in 𝑉𝑆 should reflect in N-values as well, especially because of

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 50

uncertainly in hammer energy, as discussed. Moreover, the fluctuations observed at a few depth,
i.e., extreme low values of 𝑉𝑆 might not be real, as such soft layer should have been detected in
borelog as well.
The layer boundaries are considered at the recording depths which is not often a correct
representation of field condition. This assumption may lead to velocity proiles which may have
unrealistic fluctuations. Hence, the actual layer boundaries which are derived from borelog need
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

to be considered. However, the borelogs are often made with visual inspection of soil samples
obtained from SPT. These may not reflect actual layer boundaries often as these inspections are
subjected to human error. Inclusion of soil layer classification based on lab tests can be helpful in
such cases. Hence, including actual determined layer boundaries from borelog and lab tests can
be considered as a future scope for this study.

Figure 6. (a) Soil profile from borelog (b) Vertical arrival time for the downhole test at DH
2, (c) 𝑽𝑺 from direct method and Snell's Raypath method, (d) 𝑽𝑺 from Integrated method
and MASW along with SPT N-values

CONCLUSION

Downhole tests were conducted at two locations with silty deposits underlain by
sedimentary rocks in East India. The boreholes for the downhole test were prepared based on
ASTM 7400 recommendations. Borelog obtained during borehole drilling was utilized in
determining the depth of analysis using Snell's Law. 𝑉𝑆 profiles for the two locations were
generated using an integration of the direct method and Snell's Raypath method. This
integrated approach gave a better 𝑉𝑆 profile as compared to using a single method alone.
However, it still involves the individual generation of two profiles and later integrates them
based on shallow and deep soil profiles obtained from the borelog. Moreover, some
fluctuations are observed in the profiles obtained from Snell's Raypath method which may
not well represent actual field condition because of extreme low or high 𝑉𝑆 values. This

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 51

limitation can be addressed in extended studies where methods can be developed for
simultaneously calculating the two profiles and integrating them based on subsurface layer
thickness and properties obtained from lab tests.

ACKNOWLEDGEMENTS
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

The authors thank Research Assistants at IISc Mr. Ravinesh Kumar, Mr. Siriwanth Kumar,
and Ms. Divyashree Varadaraj for their valuable assistance during field data acquisition and data
processing. The authors thank the Dam Safety (Rehabilitation) Directorate, Central Water
Commission, for funding the project entitled "Capacity Buildings in Dam Safety" under Dam
Rehabilitation and Improvement Project". Author thanks M/s. SECON Private Limited,
Bangalore for funding project "Effect of Shear Wave Velocity Calibration on Amplification of
Shallow and Deep Soil Sites."

REFERENCES

Anbazhagan, P., Kumar, A., Ingle, S. G., Jha, S. K., and Lenin, K. R. (2021). Shear Modulus
from SPT N-value with different Energy Values. Soil Dynamic snad Earthquake
Engineering, 150, 106925.
ASTM International. (2019). Standard Test Methods for Downhole Seismic Testing, ASTM
D7400-19, Feb. 2019. ASTM, West Conshohocken, Pennsylvania, United States Available:
https://www.astm.org/Standards/D7400.htm.
Bang, E. S. (2001). The Evaluation of Shear Wave Velocity Profiles Using Downhole and
Uphole Test. Masters Thesis, The Department of Civil and Environmental Engineering,
KAIST, Daejeon, Korea.
Hwang, S., Menq, F., Stokoe, K. H., Lee, R. C., and Roberts, J. N. (2018). Advanced Data
Analysis of downhole seismic records. Geotechnical Earthquake Engineering and Soil
Dynamics V.
Joh, S. H., and Mok, Y. J. (1998). Development of an Inversion Analysis Technique for
Downhole Seismic Testing and Continuous Seismic CPT. Journal of Korea Geotechnical
Society, Vol. 14, No. 3, pp. 95–108.
Kim, D. S., Bang, E. S., and Kim, W. C. (2004). Evaluation of various downhole data reduction
methods for obtaining reliable Vs profiles. Geotechnical Testing Journal, vol. 27, no. 6, pp.
585–597.
Mok, Y. J. (1987). Analytical and experimental studies of borehole seismic methods. Ph.D.
thesis, The Department of Civil Engineering, The University of Texas at Austin, Austin,
1987.
Stokoe, K. H., Hwang, S., Roberts, J. N., Menq, F. M., Keene, A. K., Lee, R. C., and Redpath, B.
B. (2017). Deep Downhole Seismic Testing Using a Hydraulically-Operated, Controlled-
Waveform Vibroseis. 16th World Conference on Earthquake Engineering, 16WCEE,
Santiago, Chile.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 52

Remote Sensing Using Satellite Derived Products to Assess Sinkhole Occurrence

Ronald J. Rizzo1 and L. Sebastian Bryson, Ph.D., P.E., D.GE, F.ASCE2


1
Dept. of Civil Engineering, Univ. of Kentucky, Lexington. Email: ronald.rizzo@uky.edu
2
Dept. of Civil Engineering, Univ. of Kentucky, Lexington. Email: sebastian.bryson@uky.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ABSTRACT

Karst subsidence or localized cover collapse of subsurface cavities (sinkholes) is a severe


hazard in the United States and can be found in all 50 states. Although occurrences of subsidence
or sinkholes are challenging to predict, karst areas are more susceptible and conducive to the
formation of cover collapse sinkholes. Seasonality and daily conditions in soil moisture, surface
temperature, and precipitation collected from satellite-based sensors and remote sensing
algorithms can be used to investigate incipient failure factors that initiate a sinkhole: in
particular, the NASA Soil Moisture Active Passive (SMAP) Level 4 root zone soil moisture, the
Integrated Multi-satellite Retrievals (IMERG) algorithm for the Global Precipitation
Measurement (GPM) mission, and the Famine Land Data Assimilation System (FLDAS) Noah
Land Surface Model. This study presents a holistic assessment of several known sinkhole
occurrences in the United States using coarse low spatial resolution data. It was observed that
comparing the spatial and temporal volumetric soil moisture from SMAP_L4 and FLDAS Noah
with the precipitation intensity from GPM_IMERG data functioned well in detecting impending
conditions at the investigated sites. This research aimed to develop a routine process flow
analysis through SMAP L4_SM, FLDAS Noah, and GPM_IMERG data and investigate the
normalized difference vegetation index (NDVI) to conduct sinkhole assessment.

INTRODUCTION

Sinkholes hazards are severe geohazards found in all 50 states of the United States and many
countries around the world. The geologic formations most associated with sinkhole occurrences
are areas of karst formations. Karst formations are characterized by soluble rocks such as
carbonate (limestone and dolomite) and evaporate (salt, gypsum, and anhydrite) rocks. Changes
in the subsurface void space in karst formations are due to the dissolution of the geologic
materials and the dissolution is associated with changes in the hydrological conditions (Tihansky
1999). Understanding the incipient hydrological, geologic, climatological, and geomorphological
conditions leading to the formation of sinkholes is fundamental to forming disaster risk
management and mitigation strategies. Thus, it is theorized that these conditional factors can be
directly combined to indicate potential sinkhole activity (Nam et al., 2020). Some incipient
conditional factors leading to the formation of sinkholes can be obtained from satellite-based
earth and precipitation data. For example, meteorological conditions (e.g., precipitation and
surface temperature data) can be extracted from the National Aeronautics and Space
Administration (NASA) Global Precipitation Measurement (GPM) mission with Integrated
Multi-satellite Retrievals (IMERG) for GPM. Satellite-measured volumetric soil moisture can be
acquired from the NASA Soil Moisture Active Passive (SMAP) satellite mission. The NASA
SMAP satellite mission is an orbiting observatory with the primary purpose of providing global
mapping of soil moisture and landscape freeze/thaw state (Entekhabi et al., 2010; O'Neill et al.,

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 53

2010). In addition, soil moisture content can be collected from the Famine Land Data
Assimilation System (FLDAS) Noah Land Surface Model, adapted from the Land Information
System (LIS7), and changes in vegetation in areas of collapse can be assessed from the
Normalized Difference Vegetation Index (NDVI).
This study investigated the viability of using near-real-time high-quality, low-latency spatial
and temporal volumetric water content and precipitation data obtained from space-based
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

platforms to assess the potential conditions that proceed sinkhole formation in karst
environments.

MATERIAL AND METHODS

Soil Moisture – Surface and Root Zone Soil Moisture

Root-zone soil moisture is commonly simulated through land surface models that physically
relate to the surface (0-5 cm) and root-zone (0-200 cm) layers. The SMAP Level 4 Surface and
Root Zone Soil Moisture (SMAP L4_SM) data products were used for this research. The SMAP
L4_SM product is a model-derived value-added product obtained by merging SMAP
observations with estimates from a land surface model in a data assimilation system. The land
surface model component of the assimilation system is driven by observation-based
meteorological forcing data, including precipitation, which is the most important driver for soil
moisture. The model-derived product produced 3-hourly estimates of surface and root zone soil
moisture (to a depth of 100 cm) at a 9 km gridded resolution with a data availability latency of 7
to 14 days (Chan et al., 2016). Also used in the research was the (FLDAS) Noah Land Surface
Model. The FLDAS Noah dataset contains a series of 28 land surface parameters in a 0.1-degree
spatial resolution. Estimates are produced monthly (McNally 2019).

Observed Satellite Precipitation

IMERG (Integrated Multi-Satellite Retrievals for Global Precipitation Mission) Level-3


precipitation dataset was used for understanding the indirect water content near the identified
sinkholes used in the study. Level-3 results represent the final daily accumulated precipitation
from near-real-time global multi-satellite estimates. Full spatial resolution for Level-3 products
from IMERG is 0.1-degree latitude by 0.1-degree longitude, and the temporal resolution can vary
between 30 minutes and monthly. Therefore, the latency from satellite acquisition of Level-3 can
range from 4 hours with an “early run”, 12 hours with a “late run”, and 3.5 months with a “final
run” (Huffman et al., 2015).

Normalized Difference Vegetation Index (NDVI)

The Normalized Difference Vegetation Index (NDVI) is an index derived from the near-
infrared wavelength reflections from the cell structure of leaves to provide consistent spatial and
temporal comparisons of vegetation greenness. The vegetation indices are collected from the
Moderate Resolution Imaging Spectroradiometer (MODIS) instrument aboard the multi-national
NASA scientific satellite TERRA, produced at 16-day intervals at a 500 m spatial resolution
(Kondoh and Higuchi, 2001). The mathematical formula for deriving the NDVI of plant growth
on earth is written as (Seeyan et al., 2014):

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 54

NIR − VIS
NDVI = (1)
NIR + VIS

where NIR = near-infrared spectral wavelength; VIS = land surface reflectance in the visible
wavelength. Calculations of NDVI result in values ranging from -1 to +1, where 0 indicates no
vegetation and closer to +1 (0.8 – 0.9) indicates the highest possible vegetation.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Study Area

Sinkholes defined in this study were selected from multiple locations in the US and are
shown in Figure 1.

a.) Louisville Kentucky Sinkhole

Sinkhole Area

b.) Branson Missouri Sinkhole


Sinkhole Area

c.) Land O’Lakes Florida Sinkhole


Sinkhole Area

Figure 1. Images of three known cover collapse sinkholes: (a) Louisville Zoo Sinkhole that
occurred in Louisville, Kentucky, on March 06, 2019; (b) the Top Rock golf course sinkhole
was discovered in Branson, Missouri, on May 5, 2015; and (c) Land O’Lakes sinkhole
occurred in Land O’Lakes, Florida on July 13, 2017. (Images developed using Google
Earth)

The Louisville Zoo sinkhole is located at Louisville, Kentucky, at latitude 38 12’12.76” N
and longitude 8542’15.47” W (Figure 1a), estimated size to be 54 meters wide, 82 meters long,

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 55

and 15 meters deep. The sinkhole occurred on March 06, 2019. The subsurface and underlying
rock near the study area are part of the Bluegrass Region, characterized by fossiliferous
limestone, shale, and dolomite. The Kentucky Geological Survey identifies the region as a karst
landscape subject to developing features like sinkholes, caves, and springs.
The Branson sinkhole is located at the Top Rock golf course in Branson, Missouri, at latitude
36 32’28.18” N and longitude 9314’58.24” W (Figure 1b). The initial size of the sinkhole was
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

22 meters wide and 11 meters deep in some areas. The sinkhole appeared on May 5, 2015. The
area was later discovered to be located over an extensive karst cavern system.
The Land O’Lakes sinkhole is in Florida at latitude 28 12’12.784” N and longitude
8227’19.80” W (Figure 1c) with the dimensions of 42 meters in width, 50 meters in length, and
12 meters deep. The sinkhole was discovered on July 13, 2017. According to the Florida
Speleological Society, the soil stratigraphy near the sinkhole is part of the Ocala Uplift
limestone. The limestone bedrock may be very close to the surface in some areas, within 0 to 15
meters, and exhibits extensive karstification.

RESULTS AND DISCUSSION

Volumetric soil moisture from SMAP and FLDAS NOAH missions

Radiometer instrumentation onboard the NASA SMAP mission provided spatially in-situ soil
moisture measurements at the top 200 cm (vertical average) below the surface at the native grid-
cell scale resolution of 9-km. Figure 2 shows the volumetric root zone soil moisture for each site
for one-year and three-year periods retrieved before and after the reported sinkhole failure. The
three-year investigation period was selected to provide consistent analogy-based availability of
the various remote satellite-based data.
The volumetric soil moisture level for the Louisville sinkhole was at a seasonal low of 0.256
(m /m3) during September 2018. By February 24, 2019, the moisture content reached a 3-year
3

high of 0.432 (m3/m3), and failure occurred ten days later, as shown in Figures 2a and 2b. The
FLDAS NOAH data in Figure 2c and Figure 2d for the Branson sinkhole followed a similar
trend with the seasonal low moisture content of 0.236 (m3/m3) in September 2014 and failed after
the moisture content reached a high value of 0.37 (m3/m3) (Figure 2c and 2d). The Land O’ Lake
sinkhole data in Figure 2e and Figure 2f showed a dry season that started in September 2016
with a 3-year high of 0.18 (m3/m3) and steadily declined until May 2017 to a 3-year low of 0.033
(m3/m3). The sinkhole occurred following a sharp increase in moisture content to 0.118 (m3/m3)
in July 2017.
Figure 2 shows that all three cover collapse sinkholes occurred during the wetting season,
with a short dry period between precipitation events. Nevertheless, it is essential to note that the
precipitation intensity rate at the time of sinkhole failure was not the highest for any three sites.
The Louisville and Branson locations experienced the maximum precipitation levels at the time
of sinkhole failure.

Annual and cumulative GPM (IMERG) precipitation data

Previous studies indicate that cover-collapse and cover-suffusion development may be


related to hydrological/hydrogeological conditions such as precipitation and groundwater
recharge. Figure 3 shows the IMERG satellite data for the three sinkholes. According to Figure

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 56

3a, the Louisville sinkhole site experienced rainfall from February 5 through March 6, 2015,
with a total accumulation of 0.34 m. This wetting period was followed by an 11-day drying
period during which the sinkhole occurred (Figure 3a).
0.5 0.45
a.) SMAP-SPL4SMGP_1 YR b.) SMAP_SPL4SMGP_3 YRS
Sinkhole occurence
Volumetric Soil Moisture (m3/m3)

Sinkhole occurence

Volumetric Soil Moisture (m3/m3)


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

0.45
0.4

0.4
0.35
0.35
0.3
0.3

0.25
0.25

0.2 0.2

0.4 0.4
c.) NCALDAS_NOAH_1 YR d.) NCALDAS_NOAH_3 YRS
Sinkhole occurrence Volumetric Soil Moisture (m3/m3) Sinkhole occurrence
Volumetric Soil Moisture (m3/m3)

0.35
0.35

0.3
0.3
0.25

0.25
0.2

0.2 0.15

0.15 0.1

0.2 0.22
SMAP_SPL4SMGP_1 YR SMAP_SPL4SMGP_3 YRS
e.) f.)
0.18 0.2 Sinkhole occurrence
Volumetric Soil Moisture (m3/m3)

Sinkhole occurrence
Volumetric Soil Moisture (m3/m3)

0.18
0.16
0.16
0.14
0.14
0.12
0.12
0.1
0.1
0.08
0.08
0.06 0.06
0.04 0.04

0.02 0.02

Figure 2. Graphs of annual and three-year root zone soil moisture for the three study sites
(a) Louisville 1-year; (b) Louisville 3-year; (c) Branson 1-year; (c) Branson 3-year; (e)
Land O’Lakes 1-year; (f) Land O’Lakes 3-year used in this research. Sinkhole events are
indicated as red lines on each graph.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 57

120 3000 200 3000

Precipitation Cumulative - Late Run (mm/day)


GPM_IMERGDL GPM_IMERGDL

Precipitation Cumulative - Late Run (mm/day)


a.) Sinkhole occurrence
GPM_IMERG_CUMULATIVE 180 b.) Sinkhole occurrence
GPM_IMERG_CUMULATIVE
Precipitation -LateRun (mm/day) 100 2500 2500

Precipitation -LateRun (mm/day)


160
140
80 2000 2000
120
60 1500 100 1500
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

80
40 1000 1000
60

20 500 40
500
20
0 0 0 0

90 4500
GPM_IMERGHHL

Precipitation Cumulative - Late Run (mm/day)


c.) Sinkhole occurence
80 GPM_IMERG_CUMULATIVE
4000
Precipitation -LateRun (mm/day)

70 3500

60 3000

50 2500

40 2000

30 1500

20 1000

10 500

0 0

Figure 3. The Integrated Multi-Satellite Retrievals for GPM (IMERG) graphs provide
daily precipitation and cumulative precipitation for (a) Louisville Sinkhole; (b) Branson
Sinkhole; (c) Land O’Lakes Sinkhole. Sinkhole events are indicated as red lines on each
graph.

The Branson site (Figure 3b) received a cumulative rainfall of 1.2 m during the 67 days
spanning from March through May 2015, with a total dry time equaling 19 days. The data shows
two significant rainfall events at the start of this wetting period: (1) on March 8 through March
17, 2015, with a total of 0.372 m rainfall, and (2) from March 22 – to March 27, 2015, with an
additional 0.351 m precipitation (Figure 3b). The Land O’ Lake’s area consistently received 64
days of rainfall totaling 1.52 m from March 2017 until the sinkhole occurrence on July 13, 2017,
with a combined drying time of 17 days (Figure 3c).
Table 1 shows the cumulative IMERG satellite precipitation data from the peak of one
wetting season to the peak of another.
In Table 1, it was observed that both Louisville and Branson received the highest
precipitation values during the year of sinkhole detection. Conversely, Land O’Lakes had the
lowest precipitation value for the period before the sinkhole occurrence.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 58

Table 1. Cumulative precipitation for each of the 3 years was obtained from GPM
(IMERG) from the peak of one wetting season to the peak of the next wetting season.

Louisville Branson Land O'Lakes


Dates Cumulative Dates Cumulative Dates Cumulative
Precipitation Precipitation Precipitation
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(mm) (mm) (mm)


4/1/2016- 2102.25 6/1/2012 - 1811.64 8/1/2014- 2374.91
3/31/2017 5/31/2013 7/31/2015
4/1/2017- 2521.29 6/1/2013 - 1854.99 8/1/2015- 2276.8
3/31/2018 5/31/2014 7/31/2016
4/1/2018- 2649.66 6/1/2014 - 2232.25 8/1/2016- 2091.17
3/31/2019 5/31/2015 7/31/2017

Annual groundwater recharge from IMERG GPM Precipitation

Karst landscapes are distinctive topography that indicates the dissolution of underlying
soluble limestone rocks by surface water or groundwater. Although NASA satellite missions
provide a high-resolution global mapping of soil moisture and precipitation, groundwater flow
can be challenging to determine from remote sensing (Figures 2 and 3). Examining an empirical
method of calculating groundwater recharge from the GPM IMERG precipitation may reveal an
increase in subsurface water, which could indirectly indicate infiltration. An empirical approach
used to estimate the annual groundwater recharge rate for each of the three sinkhole areas
utilizing the formula in Equation 2 (Kumar 2000) and cumulative precipitation from Table 1:

R = 1.35( P −14)
0.5
(2)

where R = Groundwater recharge due to precipitation during a one-year cycle (mm); P =


Cumulative precipitation from Table 1 (mm).

Table 2. Estimated annual groundwater recharge rate.

Louisville Branson Land O'Lakes


Dates Estimation of Dates Estimation of Date Estimation of
Groundwater Groundwater Groundwater
Recharge (mm) Recharge (mm) Recharge (mm)
4/1/2016- 61.69 6/1/2012 - 57.24 8/1/2014- 65.6
3/31/2017 5/31/2013 7/31/2015
4/1/2017- 67.6 6/1/2013 - 57.92 8/1/2015- 64.22
3/31/2018 5/31/2014 7/31/2016
4/1/2018- 69.31 6/1/2014 - 63.58 8/1/2016- 61.53
3/31/2019 5/31/2015 7/31/2017

Based on the estimated annual groundwater recharge rate (R) calculated in Table 2, both
Louisville and Branson indicated the highest estimated groundwater recharge from precipitation
during the period of each sinkhole occurrence. There are two reasons why this is important for

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 59

sinkhole collapse triggering. Firstly, clay soil failure due to wetting is associated with the loss of
shear strength from the increased groundwater recharge, which reduces soil suction. Decreased
soil suction equates to reduced effective stress in the soil leading to collapse (Miller et al., 2015).
Secondly, karst landscapes are distinctive topography that indicates the dissolution of underlying
soluble limestone rocks by surface water or groundwater. This observation may help explain the
sinkhole occurrences in Louisville and Branson, which are in karst landscapes with a clay
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

overburden.
On the contrary, approximately forty percent of new sinkholes in Central Florida occur when
the groundwater recharge is at seasonally low levels (Wilson and Beck 1992). The Land O’Lakes
sinkhole formation is in the Florida karst terrain, with the overburden consisting mainly of fine
sand. Granular soils are influenced by wetting through capillary cohesion, which increases soil
strength as water content increases until a level of saturation is reached. The Land O’ Lakes
findings from Table 2 indicate seasonally low groundwater, coinciding with hydrogeologic
factors for sinkhole occurrence in the region.

Figure 4. The NDVI graphs were derived from near-infrared wavelength reflections from
the cell structure of leaves during a 3-year cycle for the sinkhole sites of (a) Louisville, (b)
Branson, and (c) Land O’Lakes. Sinkhole events are indicated as red lines on each graph.

NDVI and subsurface water content

The vegetation indices in Figure 4 provide an additional indicator of the accessibility of


subsurface water at the three study sites over three years leading up to each sinkhole failure. In

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 60

addition, understanding the vegetation at the soil surface could provide further information on
the effects of plant root cover infiltration capacity, yielding less runoff. The Louisville site
recorded a 0.015 index in January 2019, with the sinkhole occurring in March 2019 at 0.2492
(Figure 4a). Branson observed the lowest index value of 0.1552 in February 2015, followed by
the highest value of 0.8258 on the day of the sinkhole occurrence in May 2015 (Figure 4b). Land
O’Lakes site experienced the second-highest index value of 0.5929 in June 2017 and the third
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

highest value of 0.5569 in May 2017, with a decline of 0.1443 between the two time periods.
The sinkhole occurred in July 2017, with an index of 0.4527(Figure 4c).
The NDVI data for the Louisville and Branson sinkholes indicated that both sites developed a
sinkhole after experiencing the lowest vegetation index within the three years of data retrieved
for each site. In addition, the vegetation indices concur with groundwater recharge calculations
in table 2, indicating water infiltration into the subsurface. Therefore, NDVI findings that all
three sites were either below average (Figures 4a) or above average (Figure 4b and 4c) seasonal
vegetation indices.

CONCLUSION

In conclusion, remote sensing data examination of volumetric soil moisture content, localized
precipitation, and vegetation indices demonstrate a correlation between sinkhole failure and soil
surface conditions in all three regions of interest. Therefore, analyzing the intensity and
cumulative precipitation data over several years could provide insight that may help determine
the soil moisture conditions before failure. Based on the findings of this study, seasonal effects
along with the temporal domain may be further investigated to improve the understanding of
sinkhole occurrences. In addition, combined effects of the current findings with changes in
geological conditions may further contribute to the knowledge of cover collapse formation.

REFERENCES

Chan, S. K., et al. (2016). “Assessment of the SMAP passive soil moisture product.” IEEE
Transactions on Geoscience and Remote Sensing, 54(8), 4994-5007.
Entekhabi, D., et al. (2010). “The soil moisture active passive (SMAP) mission.” Proceedings of
the IEEE, 98(5), 704-716.
Kondoh, A., and Higuchi, A. (2001). “Relationship between satellite‐derived spectral brightness
and evapotranspiration from a grassland.” Hydrological Processes, 15(10), 1761-1770.
Kumar, C. P. (2000). “Groundwater assessment methodology.” Technical Note, National
Institute of Hydrology, Roorkee, India.
Miller, G. A., Cerato, A. B., Hassanikhah, A., Varsei, M., Doumet, R., Bourasset, C., and Bulut,
R. (2015). “The Effects of Soil Suction on Shallow Slope Stability.” Final Report, FHWA-
OK-15-04, Oklahoma Department of Transportation, Oklahoma City, OK.
Nam, B. H., Kim, Y. J., and Youn, H. (2020). “Identification and quantitative analysis of
sinkhole contributing factors in Florida’s Karst.” Engineering Geology, 271, 105610.
McNally, A., and NASA/GSFC/HSL. FLDAS Noah land surface model L4 global monthly 0.1×
0.1 degrees (MERRA-2 and CHIRPS). NASA/GSFC/HSL, Greenbelt, MD, USA, Goddard
Earth Sciences Data and Information Services Center (GES DISC).
https://giovanni.gsfc.Nasa.gov/Giovanni/Accessed March 17, 2019.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 61

O’Neill, P., Entekhabi, D., Njoku, E., and Kellogg, K. (2010). “The NASA soil moisture active
passive (SMAP) mission: Overview.” Proc., 2010 IEEE International Geoscience and
Remote Sensing Symposium, IEEE, 3236-3239.
Seeyan, S., Merkel, B., and Abo, R. (2014). “Investigation of the relationship between
groundwater level fluctuation and vegetation cover by using NDVI for Shaqlawa Basin,
Kurdistan Region-Iraq.” Journal of Geography and Geology, 6(3), 187.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Tihansky, A. B. (1999). “Sinkholes, West-Central Florida.” Land subsidence in the United


States: US Geological Survey Circular 1182, 121-140.
Wilson, W. L., and Beck, B. F. (1992). “Hydrogeologic factors affecting new sinkhole
development in the Orlando area, Florida.” Groundwater, 30(6), 918-93.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 62

Statistical Analysis of Undrained Strength as Linear Function of Depth

Prince Turkson, Ph.D.1; and Daniel R. VandenBerge, Ph.D., P.E.2


1
Black and Veatch, Houston, TX. Email: turksonp@bv.com
2
Associate Professor, Tennessee Tech, Cookeville, TN. Email: dvandenberge@tntech.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ABSTRACT

The undrained shear strength of clay often increases in an approximately linear manner with
depth. This trend can be characterized by an undrained strength at the top of the soil layer (su0)
and a change in undrained strength with depth (). Because naturally occurring soils are rarely
homogeneous and consequently show high variability in undrained shear strength, the measured
properties will exhibit scatter about the linear trend. While many studies have examined the
variability in undrained shear strength itself, statistical data for describing undrained shear
strength as a function of soil depth is sparse to nonexistent in the geotechnical literature. In
addition, the slope and intercept parameters would be expected to show significant correlation,
which has not been investigated. The analysis presented in this paper estimates the intercept,
slope, and statistics of published data sets of undrained shear strength and generates smaller
synthesized data sets based on that information. The synthesized data is then used to estimate the
mean and standard deviation of the slope and intercept parameters from the linear trends, and to
estimate the correlation coefficient between these two parameters. A strong negative correlation
is found between su0 and ρ. The stability of a homogenous clay slope with linearly increasing
undrained shear strength is examined probabilistically. The negative correlation between su0 and
ρ substantially reduces the calculated probability of failure for the example.

INTRODUCTION

Undrained shear strengths (su) are difficult to accurately evaluate owing to the various factors
which influence undrained strength values for relatively low permeability soils. Nonetheless, the
use of undrained strength methods has proven to be sufficient in many geotechnical engineering
problems. Various field or laboratory tests can be used to obtain undrained shear strength
parameters. Some of the field tests include field vane shear and cone penetration, while some
laboratory tests include unconsolidated-undrained triaxial, unconfined compression, and direct
simple shear tests. Typically, in an undrained strength analysis, the friction angle (ϕ) is assumed
to be equal to zero for saturated conditions. The undrained shear strength of a soil mass can be
represented in a variety of ways. One common approach is to express the undrained shear
strength as a single value that is assigned to the entire soil matrix. Another method is to
characterize undrained shear strength as a linear or non-linear variation with soil depth with
some initial strength value specified at the ground surface (e.g., Skempton 1948, Griffiths and
Yu 2015). The second method is examined in this study.
Naturally occurring soils are rarely homogeneous and consequently show high variability in
their measured properties. Variability is often quantified based on existing data through
statistical analysis, and in the absence of adequate existing data, appropriate variability may be
assigned using engineering judgment (Sleep and Duncan 2014). The variability of undrained
shear strength itself has been examined by many studies (e.g., Sleep and Duncan 2014; Kulhawy

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 63

1992; Harr 1984; Lacasse and Nadim 1997; Phoon and Kulhawy 1999; Phoon et al. 1995).
However, currently little to no statistical data is available for the parameters which describe the
variation of undrained shear strength as a function of soil depth. The primary goal of this paper is
to present a method for evaluating the statistics of undrained shear strength as a linear function of
soil depth with some initial strength value specified at the top of soil layer when limited
measured data is available. The proposed approach is useful for probabilistic analyses where
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

statistics are required for the parameters, rather than simply selecting a conservative percentile.
An example is provided to illustrate the stability of a saturated clay slope with linearly increasing
undrained shear strength.

BACKGROUND

Lumb (1966) showed that most soil properties, including undrained shear strength, can be
regarded as random variables. Undrained shear strength with depth data from in situ vane test for
marine clay were provided by Lumb (1966) as part of studies to investigate variations in
strength.
Modeling undrained shear strength as a linear function of soil depth requires two parameters,
i.e., undrained strength at the top of soil layer, also referred to as intercept (su0) and the change in
undrained strength with soil depth, also referred to as slope (ρ), which are shown in Figure 1.
Statistical data for these two parameters, as well as real data from measured undrained shear
strength with soil depth is sparse to nonexistent in literature.

Figure 1. Undrained strength as a linear function of depth

In recent work by Zhu et al. (2021), the reliability of infinite slopes with linearly increasing
mean undrained strength was assessed. In their study, random undrained strength was generated
using an algorithm based on a non-stationary random field method to evaluate the influence of
undrained strength variability on the reliability of infinite slopes. Zhu et al. (2021) showed that
for an infinite spatial correlation length or “single random variable”, the probability of failure for
an infinite slope with linearly increasing mean strength is always underestimated. This
observation calls for the need to consider the influence of variability in the values of su0 and ρ

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 64

selected for geotechnical analysis. Other important applications for considering the variability of
strength increase with depth are the characterization of marine sediments (e.g., Kopf et al. 1998,
Yong and Qiang 2018), natural gas production (Yang et al. 2019), and slope stability (Griffiths
and Yu 2015).
Variability of soil properties is often represented by the coefficient of variation, COV=σ/μ,
where σ is the standard deviation and μ is the mean. Typical COV values for su range from 13 to
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

40% (Sleep and Duncan 2014; Kulhawy 1992; Harr 1984; Lacasse and Nadim 1997)
Whenever a soil property is described using a trend with more than one parameter (e.g., c’
and ’ or intercept and slope), the parameters may be statistically correlated. In other words, the
value of one parameter may affect the other parameter. Positive correlation indicates that the
parameters tend to increase or decrease together. A negative correlation causes one parameter to
tend to decrease when the other increases. The correlation coefficient, r, can be used to describe
this relationship as:

Cov ( x1 , x2 )
r12 = (1)
x x
1 2

where Cov is covariance, x1 and x2 are correlated variables.

METHODOLOGY

Ideally, the statistics of su0 and ρ could be determined based on trends from many locations
(i.e., boreholes or cone penetration tests) in the same deposit. However, in many cases, sufficient
data is not available to obtain such trends at enough locations to develop meaningful statistics. In
absence of real data at multiple locations on the same site, this study explores a method that uses
synthetic data to estimate statistics for su0 and ρ.
In order to determine statistical values to use for su0 and ρ, undrained shear strength with
depth data from four studies in the literature were examined. The data are summarized in Figure
2. The statistics of each data set were used to generate synthetic data sets, which were used to
estimate COV values for su0 and ρ and to evaluate the magnitude of the correlation coefficients.
The procedure used is summarized in the following paragraphs.
For each of the four data sets, the mean su0 and ρ were determined by fitting a linear trendline
to the measured data as shown in Figure 2. Rather than attempt to determine the variation of the
trends directly from the sparse data sets provided by most of the sources, a constant COV of 20%
was selected for undrained shear strength. The selection of COV equal 20% was based on typical
range of values reported by Sleep and Duncan (2014), Kulhawy (1992), Harr (1984), and
Lacasse and Nadim (1997). In addition to the 20% COV for su, data from Yang et al. (2019) was
re-evaluated using an assumed COV of 10% for comparison. Lines indicating a plus or minus
one standard deviation range about the mean are shown on Figure 2.
Synthetic data sets were generated using the four data sets as basis. This simulates conditions
in which less data is available to determine the trend in su with depth. Within eight equally
spaced intervals, depths were selected randomly selected from a uniform distribution. The mean
trendline was used to calculate the mean undrained shear strength for each randomly selected
depth. Subsequently, for each iteration, random values of undrained strength were generated
using the mean su for each depth and COV = 20%. A normal distribution was assumed for the
undrained shear strength about the mean trend. Using the eight synthetic data points, a trendline

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 65

was fit to determine values of su0 and ρ for that iteration. Figure 3 shows an example of
undrained strength data simulation based on actual data to obtain COV values for su0 and ρ. Ten
thousand sets of synthetic data were generated for each of the four data sets from which the mean
 (), standard deviation of  (), mean su0 (su0), and standard deviation of su0 (su0) were
determined.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Undrained strength versus depth data from four sources

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 66
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Undrained shear strength simulation procedure

The methodology for determining the statistics of su0 and ρ is summarized as follows:
1. Determine the linear trend of the undrained shear strength vs. depth.
2. Determine from the data (or estimate) the COV of su with respect to the trend.
3. Generate a synthetic data set.
a. Randomly select depths for the synthetic data from regular intervals, each with a
uniform distribution. Use a number of intervals (eight was used for this study)
representative of the actual testing interval that is being modeled.
b. At each depth, calculate the mean shear strength from the trend (Step 1) and the
standard deviation from the COV of su (Step 2).
c. At each depth, generate the simulated shear strength from the trend and standard
deviation of su.
4. Fit a linear trend to the synthetic data points and record the slope and intercept.
5. Repeat Steps 3 and 4 a large number of times (10,000 iterations were used for this study).
The mean slope and intercept should tend back to the values from Step 1.
6. Calculate the values of , , su0, su0, and r12.

RESULTS AND DISCUSSION

The statistical parameters for the slope and intercept based on the simulated data are
summarized in Table 1. Note that the mean values of intercept and slope are slightly different
from those in Figure 2 because they are derived as statistics from the 10,000 synthetic data sets
generated for each soil. The COV values were estimated from these values. The covariance was
also determined between su0 and ρ, which allowed the correlation coefficients to be determined.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 67

Table 1. Statistical values for intercept and slope of the undrained strength relationships

Parameters Derived from Synthetic Data Sets


COV
Data Source μ for μ for σ for σ for COV for Correlation
for
intercept slope intercept slope intercept Coefficient,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

slope
(kPa) (kPa/m) (kPa) (kPa/m) (%) r12
(%)

Lumb (1966) 1.91 0.99 0.74 0.17 39 17 -0.84

Kopf et al.
13.7 0.54 1.95 0.31 14 58 -0.84
(1998)
Yong and
1.32 1.45 0.95 0.22 72 15 -0.86
Qiang (2019)
Yang et al.
(2019) 2.18 2.30 1.49 0.35 68 15 -0.86
COV(su)=20%
Yang et al.
(2019) 2.20 2.30 0.75 0.17 34 8 -0.86
COV(su)=10%

Based on these four data sets, the COV for su0 tends to be very high when the mean value of
su0 is low and the data has substantial dispersion. Because the mean is in the denominator, the
COV is very sensitive to small changes in the mean. The relatively small mean su0 and large
standard deviation su0 values for the data from Yong and Qiang (2019) and Yang et al. (2019)
resulted in relatively large values of COV. Re-evaluation of the Yang et al. (2019) data using an
assumed COV of 10% resulted in decreased simulated COV values but the correlation coefficient
was about the same. Considering the typical range of COV for undrained strength as a guidance
(13% to 40%), COV values for su0 and ρ for the data from Kopf et al. (1998) were selected for
use in the slope stability example presented later in the paper.
A correlation between su0 and ρ is expected and can be observed in the synthetic data.
Covariance values were computed from all 10,000 iterations for su0 and ρ for each data source,
and a correlation coefficient was calculated based on the covariance and standard deviation
values for su0 and ρ. The correlation coefficient from the four data sources ranged from -0.84 to -
0.86. An average correlation coefficient of -0.85 was calculated from the four data sources. This
implies that as the slope () or change in undrained shear strength with depth decreases, the
undrained strength at the top of the soil layer (su0) increases, and vice-versa. The large
correlation coefficients indicate a substantial dependence between the slope and intercept of the
linearly increasing undrained shear strength model. These two parameters cannot be assumed to
vary independently from each other.

APPLICATION TO SLOPE STABILITY ANALYSIS

The simple slope with inclination, X= cot(θ) as shown in Figure 4, was used as an example to
illustrate the use of the statistics calculated in this study.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 68
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Clay slope geometry

The factor of safety for a homogeneous undrained slope with linearly increasing strength was
presented by Griffiths and Yu (2015) as:


F=N (2)

where N is the stability number found from their slope stability charts or MATLAB solution and
 is the total unit weight of the soil.
The stability number depends on the strength gradient, M, defined as:

su 0
M= (3)
H

where H is the slope height. VandenBerge and McGuire (2017) developed a response surface
equation for N based on Griffith and Yu’s solution in which

N = 2.513 + 1.801X − 0.045 X 2 + 4.631M + 0.489M 2 + 1.599 XM − 0.196 X 2 M


(4)
− 0.441XM 2 + 0.057 X 2 M 2 + 

where  is a normally distributed error term with standard deviation of 0.03


Figure 4 summarizes the parameters for the clay slope and the mean undrained strength
parameters, which were adopted from Kopf et al. (1998). The probabilities of slope failure (PF)
were calculated using the Taylor series first-order second moment (FOSM) approach for two
cases: (i) correlation coefficient not considered between su0 and ρ and (ii) correlation coefficient
equal to -0.84 considered between su0 and ρ. The assumed COV for the remaining variables were
10% for H, 5.88% for γ, and 5% for θ. All the variables were assumed to be normally distributed.
The most likely value (MLV) for factor of safety for Cases (i) and (ii) is 1.205.
The factors of safety (F) calculated using the VandenBerge and McGuire response surface
are presented in Table 2. Note that factors of safety are reported to three decimal places for
calculation purposes only. However, this level of precision should not be used for final reported
values of F.
For Case i (no correlation), the standard deviation of F is approximated by:

F   ( F 2)
2
(5)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 69

where F is the difference in F from varying the variable up and down one standard deviation
about the mean. For the example, F is about 0.37, and COV(F) is 31%. Assuming that the factor
of safety is normally distributed, the reliability index is about 0.48 and the probability of failure
is about 32% for Case i.

Table 2. Taylor Series FOSM probabilistic analysis – Cases i and ii


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Variable Factor of safety ΔF (ΔF/2)2


H = 5.5 m (+1 SD) 1.136
-0.149 5.55×10-3
H = 4.5 m (-1 SD) 1.285
su0 = 15.62 kPa (+1 SD) 1.305
0.208 1.08×10-2
su0 = 11.78 kPa (-1 SD) 1.097
ρ = 0.853 kPa/m (+1 SD) 1.424
0.676 1.14×10-1
ρ = 0.227 kPa/m (-1 SD) 0.747
 = 18 kN/m3 (+1 SD) 1.138
-0.142 5.06×10-3
 = 16 kN/m (-1 SD)
3
1.281
θ = 21 deg (+1 SD) 1.201
-0.009 2.11×10-5
θ = 19 deg (-1 SD) 1.211

For Case ii (correlation between su0 and ), the standard deviation of F is approximated by:

F   ( F 2) + ( r F F ) / 2
2
12 su 0
(6)

For the example, F for Case ii is about 0.28, and COV(F) is 23%. Assuming that the factor
of safety is normally distributed, the reliability index is about 0.71 and the probability of failure
is about 24% for Case ii.
The statistics for the two cases considered are summarized in Table 3. Reliability of the
undrained homogeneous slope increases when the correlation between su0 and ρ is considered as
compared to Case i where su0 and ρ are assumed to be mutually independent of each other. In the
example presented, the correlation coefficient between su0 and ρ resulted in a 9% lower
probability of failure for Case (ii) compared to Case (i). Ignoring the correlation between the
slope and intercept of the undrained shear strength trend in a stability analysis can lead to an
overconservative result because of the strong negative correlation between su0 and ρ.

Table 3. Probabilities of failure for the example

Case Approx. COV of F Reliability index,  Approx. PF


i 31% 0.48 32%
ii 23% 0.71 24%

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 70

CONCLUSION

Uncertainty associated with homogenous clay soils with strength as a linear function of depth
was assessed in this study. Methods to statistically quantify this uncertainty for the su0 and ρ
using smaller synthesized data sets based on limited measured data have been presented. The
methods and findings from this study illustrate a method for estimating the mean, standard
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

deviation, and COV for parameters of the linearly increasing undrained shear strength model
when limited field testing data is available. A constant COV of 20% was assumed for su about the
mean trend for most of the simulated data sets. The magnitude and standard deviation and COV
of su0 and ρ will be influenced by this assumption, and the reported values are specific to the
conditions analyzed.
Statistical analysis of the simulated data shows a strong negative correlation (r12 = -0.84 to -
0.86) between su0 and ρ for the soils considered. The calculated correlation coefficients may be
affected by the assumptions of the simulations (e.g., COV of su, number of simulated points,
vertical distribution of simulated points). These effects warrant further study before further
conclusions can be made about the typical magnitude of correlation between su0 and ρ.
The two cases of the example illustrate the significant influence of the correlation between
su0 and ρ, at least in the case of slope stability, on the calculated likelihood of slope failure. For
this reason, su0 and ρ for any set of undrained shear strength data cannot be treated as mutually
independent of each other in probabilistic evaluation of stability. The Taylor series probabilistic
analysis produced lower probability of failure when correlation between su0 and ρ was
considered. The effect of simulated data generated within larger number of equally spaced
intervals may be investigated in future studies. More data points will likely reduce the variability
in su0 and ρ but may not affect the correlation coefficient.

REFERENCES

Griffiths, D. V., and Yu, X. (2015). “Another look at the stability of slopes with linearly
increasing undrained strength.” Géotechnique, 65(10), 824-830.
Harr, M. E. (1984). Reliability-based design in Civil Engineering, Henry M. Shaw Lecture, North
Carolina State University, Raleigh, N.C.
Kopf, A., Clennell, M. B., and Flecker, R. (1998). “Relationship between the variation of
undrained shear strength, organic carbon content, and the origin and frequency of enigmatic
normal faults in fine-grained sediments from advanced piston cores from the Eastern
Mediterranean.” Proc., Ocean Drilling Program, 160 Scientific Results. Ocean Drilling
Program, Vol. 160, 645-661.
Kulhawy, F. H. (1992). On the evaluation of soil properties. ASCE Geotech. Spec. Publ. No. 31,
95–115.
Lacasse, S., and Nadim, F. (1997). “Uncertainties in characterizing soil properties.” Norwegian
Geotechnical Institute, 201, 49-75.
Lumb, P. (1966). “The variability of natural soils.” Canadian Geotechnical Journal, 3(2), 74-97.
Phoon, K.-K., and Kulhawy, F. H. (1999). “Evaluation of geotechnical property variability.”
Canadian Geotechnical Journal, 36(4), 625–639.
Phoon, K.-K., Kulhawy, F. H., and Grigoriu, M. D. (1995). Reliability-based design of
foundations for transmission line structures. Report TR-105000, Electric Power Research
Institute, Palo Alto, CA.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 71

Skempton, A. W. (1948). “A study of the geotechnical properties of some post-glacial clays.”


Geotechnique, 1(1), 1-16.
Sleep, M. D., and Duncan, J. M. (2014). Manual for Geotechnical Engineering Reliability
Calculations. Second edition. Virginia Tech, Blacksburg, Virginia.
Yang, S. L., Lunne, T., Andersen, K. H., D’lgnazio, M., and Yetginer, G. (2019). “Undrained
shear strength of marine clays based on CPTU data and SHANSEP parameters.” Proc. XVII
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ECSMGE-2019, Iceland.
Yong, B., and Qiang, B. (2018). Subsea engineering handbook. Second edition. Gulf
Professional Publishing. Oxford, United Kingdom.
Zhu, D., Xia, L., Griffiths, D. V., and Fenton, G. A. (2021). “Reliability analysis of infinite
slopes with linearly increasing mean undrained strength.” Computers and Geotechnics, 140,
104442.
VandenBerge, D., and McGuire, M. (2017). “Response surfaces for probabilistic analyses of
slope stability.” Proc., 19th International Conference on Soil Mechanics and Geotechnical
Engineering, Seoul, South Korea.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 72

Evaluation of the Maximum-Likelihood Estimates of Site Fundamental Frequencies for a


Subset of KiK-Net Strong Motion Data

Mohammad Yazdi1; Ramin Motamed2; and John G. Anderson3


1
Ph.D. Candidate, Dept. of Civil and Environmental Engineering, Univ. of Nevada Reno, Reno,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

NV. Email: m_yazdi@nevada.unr.edu


2
Associate Professor, Dept. of Civil and Environmental Engineering, Univ. of Nevada Reno,
Reno, NV. Email: motamed@unr.edu
3
Professor Emeritus, Nevada Seismological Laboratory, Univ. of Nevada Reno, Reno, NV.
Email: jga@unr.edu

ABSTRACT

In ground motion prediction equations (GMPEs), site characteristics are usually addressed by
the shear-wave velocity of top 30 m (VS30) or depth to 1.0 km/s shear-wave velocity isosurface
(Z1.0); however, these parameters are either inferred or limited to the shallow depths. Therefore,
using other site proxies such as site fundamental frequency is useful. In this study, a data set
from the KiK-net strong ground motion network is used to estimate the maximum-likelihood
estimate of site fundamental frequency (𝑓𝑚𝑙 ) and its corresponding amplitude (𝐴𝑚𝑙 ) using the
horizontal-to-vertical spectral ratio (HVSR) of surface ground-motions. First, the recorded
ground motions with PGA ≥ 0.1% g and Mw > 3 are selected and processed for 100 stations.
Second, the HVSR is calculated for each recorded event at each site. Third, 𝑓𝑚𝑙 and 𝐴𝑚𝑙 are
estimated for each site using previously developed automated methodologies by the authors.
These automated methodologies provide the uncertainties associated with 𝑓𝑚𝑙 and 𝐴𝑚𝑙 estimates,
whereas the conventional HVSR methods yield one value. Lastly, the results obtained in this
study are compared with a recent independent study on a similar data set. The results of
automated methods are very consistent with those of the independent study. According to the
results, using 𝑓𝑚𝑙 can be referred to using the first peak frequency (𝑓0 ) and highest peak
frequency (𝑓𝑝𝑒𝑎𝑘 ) in the HVSR curve. In addition, the findings show that the Fourier amplitude
spectrum and 5%-damped pseudo-spectral acceleration have fairly comparable performance in
estimating site fundamental frequency and its amplitude.

INTRODUCTION

Within the past decades, a wide array of research has been carried out on studying ground
motion models (GMMs). The GMMs are a set of equations for predicting the contribution of
source, path, and site effects in the observed ground motions intensity measures. Within GMMs,
various proxies can be used for predicting site effect some of which are VSZ (time-averaged shear
wave of the top Z meters, Z=5, 10, 20, and 30 m), ZX (depth at which the shear-wave velocity is
X km/s, X=0.8, 1.0, 1.5 and 2.5 km/s) and site fundamental frequency. Among these proxies,
VSZ is limited to shallow depths, and ZX is usually inferred from geophysical methods. On the
other hand, site fundamental frequency includes the properties of the underlying deep soil layers
and can help obtain site amplification.
Methods for estimating the site fundamental frequency can be categorized into two groups:
(a) using a nearby rock site as a reference (Bard 1994), and (b) using horizontal-to-vertical

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 73

spectral ratio (HVSR), which was first introduced by Nakamura (1989). The HVSR approach
assumes that the horizontal ground motion is amplified at the site fundamental frequency due to
the resonance effect, while the vertical motion does not show a significant amplification at that
frequency. Therefore, a peak can happen at the natural frequency of the site. This frequency is
assumed to be representing the site's fundamental frequency (Bonilla et al. 2002; Ghofrani et al.
2013; Kawase et al. 2011). There are two methods for obtaining site fundamental frequency from
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

HVSR. The first method is using microtremor measurements and obtaining the HVSR of noise
measurements (e.g. Nakamura, 1989). The second method is utilizing recorded surface ground
motions and computing the earthquake HVSR (e.g., Hassani and Atkinson 2016a; Lermo and
Chavez-Garcia 1993; Zhu et al. 2021). Using this method, one can estimate the site's
fundamental frequency in a fast and inexpensive way.
Significant impedance contrasts cause peaks in the HVSR curve of a site, and sites without
any peak can be rock sites or sites with gradual changes in the shear-wave velocity profile. Kwak
et al. (2017) provided a method to parameterize the peak of the site’s average HVSR curve. In
that study the first peak frequency (𝑓0 ) in the site’s HVSR curve is assumed to be the site
fundamental frequency. There are some other studies (Hassani and Atkinson 2016b) that use the
highest peak frequency (𝑓𝑝𝑒𝑎𝑘 ) rather than 𝑓0 to estimate the site fundamental frequency. Zhu et
al. (2020) compared the performance of 𝑓0 against 𝑓𝑝𝑒𝑎𝑘 in predicting the local site effect and
they concluded that although both 𝑓0 and 𝑓𝑝𝑒𝑎𝑘 are useful, 𝑓𝑝𝑒𝑎𝑘 seems to be more effective.
Yazdi et al. (2022) showed that using 𝑓𝑝𝑒𝑎𝑘 instead of 𝑓0 can sometimes bias the estimations.
Therefore, they provided a set of four automated methods, which are based on a likelihood
approach and use the events’ individual HVSR curve to estimate the site fundamental frequency
(𝑓𝑚𝑙 ) and its corresponding amplitude (𝐴𝑚𝑙 ). For the first time, the automated methods provide
the uncertainty in the estimation of site fundamental frequency using the earthquake HVSR.
Depending on the site and the HVSR curve, 𝑓𝑚𝑙 can be comparable to either 𝑓0 or 𝑓𝑝𝑒𝑎𝑘 . It
should be noted that 𝑓𝑚𝑙 can be a completely different from 𝑓0 or 𝑓𝑝𝑒𝑎𝑘 for the sites lacking a
very significant sharp peak in the average HVSR curve.
In this study, a subset of the Kiban-Kyoshin network (KiK-net), which is a strong motion
network, is utilized along with the automated methods developed by Yazdi et al. (2022) in order
to estimate the site fundamental frequency, its corresponding amplitude, and their associated
uncertainties. Then, a comparison is made between the findings of this study and the findings of
an independent study to evaluate the reliability of the results. Also, the performance of FAS- and
PSA-based calculations are assessed. The findings of this study will be used to improve the
performance of GMMs in the future.

DATASET

KiK-net is a strong motion network with 699 stations all over Japan. Although KiK-net has
recordings of surface and downhole seismograms, only the surface ground motions are processed
and used in this study. In this paper, all the recordings from 1997 to November 2021 with PGA ≥
0.001g and Mw > 3 for 100 stations are selected and processed. Consequently, 22,551 ground
motions recorded at 100 stations are included in the final dataset. Figure 1a presents the location
of the stations and Figure 1b. illustrates the magnitude-distance distribution of the data.
In addition, the findings of Zhu et al. (2021) are used for comparison purposes in this study.
That study provided the estimation of site fundamental frequency and its corresponding
amplitude obtained from the horizontal-to-vertical spectral ratio of earthquake ground motions.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 74
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. (a) Location of the selected KiK-net stations (b) Distribution of moment
magnitude (Mw) against distance

METHODOLOGY

Processing ground-motions

After finalizing the selected dataset, the procedure implemented by the NGA-West2 project
(Ancheta et al. 2014) is followed to process the recordings. This process includes tapering, zero-
padding, baseline correction, and bandpass filtering. Then, the Fourier amplitude spectrum
(FAS) and 5% damped pseudo-spectral acceleration (PSA) of the two horizontal motions and a
vertical one are obtained. Konno and Ohmachi's (1998) method with b=40 is used to minimize
the effect of smoothing on the shape of FAS (Zogh et al, 2021). Also, signal-to-noise ratio (SNR)
≥ 3 is utilized for obtaining the useable frequency range (lowest and highest useable frequency)
of the spectra.

HVSR method calculation

First, two horizontal components (e.g. North-South and East-West components) are
combined using the RotD50 method (Boore 2010) as it is orientation independent. Then, the
horizontal-to-vertical spectral ratio (HVSR) is calculated using Eq. 1.

∑𝑁𝑠
𝑗=1 log 𝐻𝑉𝑆𝑅𝑗,𝑠
log ̅̅̅̅̅̅̅̅̅
𝐻𝑉𝑆𝑅𝑠 = (1)
𝑁𝑠

In this equation, log 𝐻𝑉𝑆𝑅𝑗,𝑠 is the HVSR curve of the jth recorded event at station S; 𝑁𝑠 is
the number of recorded events at station S, and ̅̅̅̅̅̅̅̅̅
𝐻𝑉𝑆𝑅𝑠 is the average HVSR curve of station S.
Considering the findings of Yazdi et al. (2022), both FAS and PSA are used as spectral
calculation methods for computing HVSR and estimating the site fundamental frequency. Figure
2 shows the average FAS-based HVSR of the EHMH09 station with its 95% confidence interval.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 75

For sites without available ground-motion recording instrumentation, Tamhidi et al. (2021)
proposed a new approach, which constructs the ground-motion time series, FAS, and PSA at
target sites, and it can be used for estimating site fundamental frequency.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. The average HVSR curve of EHMH09 station in KiK-net. The shaded area shows
the 95% confidence interval of the average HVSR curve (solid line). The black circle shows
the median estimation of 𝒇𝒎𝒍 and 𝑨𝒎𝒍 . The error bars show the 68% confidence interval in
estimating 𝒇𝒎𝒍 and 𝑨𝒎𝒍 . The blue dashed line is the average HVSR curve obtained by Zhu
et al. (2021).

Automated methods

To estimate the site's fundamental frequency and its corresponding amplitude, the individual
HVSR curves are used as the input of four automated methods developed by Yazdi et al. (2022).
These automated methods get individual HVSR curves and the useable frequency range as the
inputs and provide the maximum-likelihood estimate of site fundamental frequency (𝑓𝑚𝑙 ), its
corresponding amplitude (𝐴𝑚𝑙 ) and their associated uncertainties as the outputs. In Method 3, the
peaks in events’ HVSR are flagged and the likelihood of occurring a peak at each sampling
frequency is computed. The frequency with maximum likelihood is selected as the site's
fundamental frequency. In this paper, the results of Automated Method 3 are presented. The
probability density function (PDF) of the estimated of 𝑓𝑚𝑙 and 𝐴𝑚𝑙 for the selected KiK-net
stations are presented in Figure 3. They tend to follow lognormal distribution. The distribution
parameters are presented in the figure.

RESULTS AND DISCUSSION

Comparison with an independent study by Zhu et al. (2021)

In this section, the estimations of Method 3 developed by Yazdi et al., (2022) are compared
with the findings of Zhu et al. (2021). Using the average HVSR curve of a site, the following

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 76

proxies are provided by Zhu et al. (2021) for KiK-net stations: the first peak frequency (𝑓0 ), its
corresponding amplitude (𝐴0 ), the highest peak frequency (𝑓𝑝𝑒𝑎𝑘 ) and its corresponding
amplitude (𝐴𝑝𝑒𝑎𝑘 ). Consequently, the 𝑓𝑚𝑙 and 𝐴𝑚𝑙 obtained from automated Method3 are
compared with 𝑓0 , 𝑓𝑝𝑒𝑎𝑘 , 𝐴0 and 𝐴𝑝𝑒𝑎𝑘 found by Zhu et al. (2021).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. The probability density function (PDF) of the estimated 𝒇𝒎𝒍 and 𝑨𝒎𝒍 for the
selected KiK-net stations. The solid line shows the lognormal distribution. 𝝁 is the mean
and 𝝈 is the standard deviation (both in log space) for each distribution.

To make a comparison, the stations are divided into three cases. In Case 1, the stations at
which 𝑓0 = 𝑓𝑝𝑒𝑎𝑘 (the first peak and the highest peak are the same) are studied. According to
Figure 4a, there is a good match between the estimations for 76 stations. Figure 5a shows an
example of a station with Case 1 condition. Note that the automated methods (Yazdi et al. 2022)
are providing uncertainty in estimations and eliminating analysts’ subjectivity. In Figure 4a,
however, there are some stations that 𝑓𝑚𝑙 is significantly different from 𝑓0 (or 𝑓𝑝𝑒𝑎𝑘 ). For
instance, station ABSH07 is shown in Figure 5b. As can been seen in the HVSR curve, there is a
very wide peak at low frequency and a wide peak at high frequency with fairly comparable
amplitudes. The automated method captures the former and Zhu et al. (2021) captured the latter.
The reason for this difference is using the maximum-likelihood approach in Method 3, whereas
the method used by Zhu et al. (2021) only uses the significant peak in the average HVSR.
In Cases 2 and 3, 8 stations with 𝑓0 ≠ 𝑓𝑝𝑒𝑎𝑘 are studied. These stations have multiple peaks
in the average HVSR curve due to multiple impedance contrast. Looking into these stations, one
can see that the automated method sometimes captures the highest peak frequency 𝑓𝑝𝑒𝑎𝑘 (Figure
4c) and sometimes captures the first peak frequency 𝑓0 (Figure 4e). As 𝑓𝑚𝑙 is the peak frequency
that is more probable to happen, and it can be close to either 𝑓0 or 𝑓𝑝𝑒𝑎𝑘 . Figures 5c and d show
sample stations for Cases 2 and 3, respectively. As was mentioned earlier, there are some cases
that 𝑓𝑚𝑙 is neither 𝑓0 nor 𝑓𝑝𝑒𝑎𝑘 (Figure 5b).
Now, the estimated amplitudes should be compared. As can be seen in Figure 4 (left panels),
the estimated amplitudes by automated methods (𝐴𝑚𝑙 ) are systematically higher than 𝐴0 or
𝐴𝑝𝑒𝑎𝑘 . There are two reasons behind this difference. The first reason is that the automated
methods use the significant peaks in the events’ individual HVSR curve and these peaks are
happening at different frequencies, whereas the conventional methods use the average HVSR of
a site and find 𝐴0 as the amplitude of the peak at only one frequency. Also, some differences in
the methods for smoothing FAS contribute to the difference between estimated amplitudes 𝐴𝑚𝑙
and 𝐴0 or 𝐴𝑝𝑒𝑎𝑘 .

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 77
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Comparison between the estimations of site fundamental frequency and its
amplitude obtained from Method 3 and the estimations of Zhu et al. (2021) study.

Figure 5. Sample stations representing different cases of comparison. The solid line shows
the average HVSR of the station, and the shaded area shows the 95% confidence interval
for HVSR. The circle shows the median estimate of 𝒇𝒎𝒍 and 𝑨𝒎𝒍 , and the error bar shows
the 68% confidence interval in estimating 𝒇𝒎𝒍 and 𝑨𝒎𝒍 . KiK-net station code is shown in
the top corner of each plot.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 78

Comparison of FAS and PSA-based Results

FAS- and PSA-based HVSR are widely used by researchers depending on the availability of
the spectra in a database. Previous studies (Stafford et al. 2017; Zhu et al. 2020) suggested that
the amplitude of PSA and the peaks in the spectrum heavily depend on the magnitude and
distance of the events recorded at one station, especially at short periods. In other words, more
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

recorded events might be required at a station if PSA is used as the spectral calculation method
compared to the FAS-based one. Therefore, there is a chance to see a significant difference
between the peak obtained from HVSRFAS and HVSRPSA. Zhu et al. (2020) showed when FAS is
used, fewer recordings are required to obtain the true resonance frequency from HVSR of ground
motion compared to the case in which PSA is used. Figure 6 compares the estimated 𝑓𝑚𝑙 and 𝐴𝑚𝑙
using FAS and PSA. For a few stations, the difference between the FAS- and PSA-based results
are significant. This difference stems from the magnitude and distance dependency of the PSA,
which affects the amplitude of the individual HVSRPSA. It can be observed that for majority of
stations, the estimations are very comparable. Consequently, both FAS and PSA-based methods
can be used considering the availability of data, and the effect of scenario dependency is
negligible when numerous recordings are available at a station. It is worth-noting that the
performance of FAS- or PSA-based results in predicting site response determines the preferred
approach.

Figure 6. Comparison between 𝒇𝒎𝒍 and 𝑨𝒎𝒍 obtained from HVSRFAS and HVSRPSA. The
solid line shows a 1:1 ratio between the compared parameters. The dashed lines represent
𝒇 𝑨
|𝐥𝐧 𝒎𝒍−𝑷𝑺𝑨 | = 𝟎. 𝟑 and |𝐥𝐧 𝒎𝒍−𝑷𝑺𝑨 | = 𝟎. 𝟑 for frequency and amplitude, respectively.
𝒇𝒎𝒍−𝑭𝑨𝑺 𝑨𝒎𝒍−𝑭𝑨𝑺

CONCLUSIONS

Various proxies can be used in ground motion models (GMMs) for predicting local site
effects. Aside from VS30 and Z1.0 which have been widely used, studies show that site
fundamental frequency is a useful complementary proxy. To estimate the site's fundamental
frequency, the horizontal-to-vertical spectral ratio (HVSR) of recorded ground motions is an
effective method. In this study, a subset of the KiK-net database with 100 stations is used for

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 79

obtaining the maximum-likelihood estimate of site fundamental frequency (𝑓𝑚𝑙 ), its


corresponding amplitude (𝐴𝑚𝑙 ) and their associated uncertainties using the HVSR method. Then,
a comparison is made between the estimations of this study (𝑓𝑚𝑙 and 𝐴𝑚𝑙 ) and those of an
independent study to assess the performance of previously developed automated methodologies.
Results show that the automated methods can reliably estimate the site's fundamental frequency,
remove the analyst’s subjectivity and quantify the uncertainty in the estimations. In addition, the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

maximum-likelihood estimate of amplitude (𝐴𝑚𝑙 ) is usually greater than the one obtained from
the average HVSR of a site since events’ individual HVSR curves are used within the automated
methods. Lastly, a comparison is made between the stations 𝑓𝑚𝑙 and 𝐴𝑚𝑙 obtained from FAS-
based and PSA-based calculations. This comparison indicates that estimations are usually more
consistent for stations with a large number of recorded ground motions. Note that these results
are based on a subset of the KiK-net database, and the whole database will be evaluated in the
future to reduce the uncertainty in GMMs by supplementing VS30 and Z1.0 with 𝑓𝑚𝑙 and 𝐴𝑚𝑙 .

DATA AND RECOURSES

The recorded ground motions from the Kiban-Kyoshin network (KiK-net) a strong motion
network (https://www.doi.org/10.17598/NIED.0004) are used in this study (last accessed
November 2022). Also, MATLAB (https://mathworks.com, last accessed May 2022) was used
for processing earthquake records, using automated methods, and making the figures.

ACKNOWLEDGEMENTS

This research was supported by the U.S. Geological Survey (USGS), Department of the
Interior, under USGS award number G19AP00100. The views and conclusions contained in this
document are those of the authors and should not be interpreted as necessarily representing the
official policies, either expressed or implied, of the U.S. Government. The authors also would
like to thank the National Research Institute of Earth Science and Disaster Resilience (NIED),
Japan for providing the KiK-net data publicly.

REFERENCES

Ancheta, T. D., R. B. Darragh, and J. P. Stewart. 2014. “NGA-West2 database.” Earthquake


Spectra 30(3): 989–1005.
Bard, P. 1994. “Effects of surface geology on ground motion: Recent results and remaining
issues.” Proc. 10Th Eur. Conf. Earthquaker Eng., 305–323.
Bonilla, L. F., J. H. Steidl, J. Gariel, and R. J. Archuleta. 2002. “Borehole Response Studies at
the Garner Valley Downhole Array, Southern Borehole Response Studies at the Garner
Valley Downhole Array, Southern California.” Bull. Seism. Soc. Am., 92 (8): 3165–3179.
Boore, D. M. 2010. “Orientation-independent, nongeometric-mean measures of seismic intensity
from two horizontal components of motion.” Bull. Seismol. Soc. Am., 100 (4): 1830–1835.
Ghofrani, H., G. M. Atkinson, and K. Goda. 2013. “Implications of the 2011 M9. 0 Tohoku
Japan earthquake for the treatment of site effects in large earthquakes.” Bull Earthquake Eng
11, 171–203.
Hassani, B., and G. M. Atkinson. 2016a. “Site-effects model for central and eastern North
America based on peak frequency.” Bull Seism. Soc Am, 106 (5): 2197–2213.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 80

Hassani, B., and G. M. Atkinson. 2016b. “Applicability of the site fundamental frequency as a
VS30 proxy for Central and Eastern North America.” Bull. Seismol. Soc. Am., 106 (2): 653–
664.
Kawase, H., F. J. Sánchez-Sesma, and S. Matsushima. 2011. “The Optimal Use of Horizontal-to-
Vertical Spectral Ratios of Earthquake Motions for Velocity Inversions Based on Diffuse-
Field Theory for Plane Waves The Optimal Use of Horizontal-to-Vertical Spectral Ratios of
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Earthquake Motions for Velocity Inversions B.” Bull Seism. Soc Am, 101 (5): 2001–2014.
Konno, K., and T. Ohmachi. 1998. “Ground-Motion Characteristics Estimated from Spectral
Ratio between Horizontal and Vertical Components of Microtremor.” Bull Seism. Soc Am, 88
(1): 228–241.
Kwak, Y., J. P. Stewart, S. J. Mandokhail, and D. Park. 2017. “Supplementing VS30 with H/V
Spectral Ratios for Predicting Site Effects.” Bull Seism. Soc Am, 107 (5): 2028–2042.
Lermo, J., and F. J. Chavez-Garcia. 1993. “Site Effect Evaluation Using Spectral Ratios with
Only One Station.” Bull Seism. Soc Am 83(5): 1574–1594.
Nakamura, Y. 1989. “A method for dynamic characteristics estimation of subsurface using
microtremor on the ground surface.” Q. Rep. RTRI, 30 (1): 25–33.
Stafford, P. J., A. Rodriguez-Marek, B. Edwards, P. P. Kruiver, and J. J. Bommer. 2017.
“Scenario dependence of linear site-effect factors for short-period response spectral
ordinates.” Bull. Seism. Soc. Am., 107 (6): 2859–2872.
Tamhidi, A., N. Kuehn, S. F. Ghahari, A. J. Rodgers, M. D. Kohler, E. Taciroglu, and Y.
Bozorgnia. 2021. Conditioned Simulation of Ground-Motion Time Series at Uninstrumented
Sites Using Gaussian Process Regression, Bull Seism. Soc Am. 112, 331–347.
Yazdi, M., R. Motamed, and J. G. Anderson. 2022. “A New Set of Automated Methodologies for
Estimating Site Fundamental Frequency and Its Uncertainty Using Horizontal-to-Vertical
Spectral Ratio Curves.” Seismol. Res. Lett., 93 (3): 1721–1736.
Zhu, C., F. Cotton, and M. Pilz. 2020. “Detecting Site Resonant Frequency Using HVSR :
Fourier versus Response Spectrum and the First versus the Highest Peak Frequency.” Bull
Seism. Soc Am, 110: 427–440.
Zhu, C., G. Weatherill, and F. Cotton. 2021. “An open-source site database of strong-motion
stations in Japan: K-NET and KiK-net (v1.0.0). Earthquake Spectra 37(3): 2126–2149.
Zogh, P., R. Motamed, and K. Ryan. 2021. “Empirical evaluation of kinematic soil-structure
interaction effects in structures with large footprints and embedment depths.” Soil Dynam
Earthq. Eng, 149: 106893. Elsevier.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 81

Simplified Calibration Procedure for Total Earth Pressure Cells

William J. Baker III, S.M.ASCE1; and Christopher L. Meehan, Ph.D., P.E., F.ASCE2
1
Graduate Student, Dept. of Civil and Environmental Engineering, Univ. of Delaware, Newark,
DE. Email: bakerwil@udel.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Professor, Dept. of Civil and Environmental Engineering, Univ. of Delaware, Newark, DE.
Email: cmeehan@udel.edu

ABSTRACT

Earth pressure cells (EPCs) are typically utilized to measure total soil and water pressures in
situ. Installed EPCs exhibit an active stress arching phenomenon in which the stress sensed by
the EPC is smaller than the stress in the soil surrounding the EPC. Unfortunately, the hydrostatic
calibration factor provided by the manufacturer cannot account for this phenomenon, resulting in
the need to independently calibrate a given EPC in the laboratory before it is installed in situ.
Though other researchers have conducted their own independent laboratory calibration of EPCs
in dry sand, the custom testing setups that have been developed for this calibration procedure
have required special apparatus fabrication, which makes these procedures difficult to replicate.
In the current study, a simplified testing apparatus and method were developed to calibrate EPCs
in dry sand, which utilize standard equipment that is commonly found in most geotechnical
testing laboratories. Three 1-MPa-range semiconductor EPCs were calibrated in dry sand at
various sand column heights ranging from 5 cm to 30 cm by subjecting them to ten loading
cycles at each column height. The goal of this study was to examine the influence of sand
column height and load cycles on the EPCs’ associated sensitivity ratio. Overall, the results
observed in the current study were consistent with results presented by previous researchers for
EPCs calibrated at the same sand column height using larger and more sophisticated testing
setups; this indicates the potential utility of the simplified calibration procedure that is described
in the current paper for independent calibration of EPCs.

INTRODUCTION

Earth pressure cells (EPCs) are used extensively in geotechnical engineering in order to
monitor earth pressures in situ. Applications of EPCs have included the monitoring of pavement
systems (Brown and Brodrick 1981, Van Duesen et al. 1992, Dawson and Little 1997),
monitoring of dynamic pressures induced during the soil compaction process (Rinehart and
Mooney 2009, Vennapusa et al. 2012), and the monitoring of dynamic lateral pressures along a
bridge abutment during the installation process of adjacent piles (Talesnick et al. 2021). The
most common types of EPCs are diaphragm cells and hydraulic cells. For diaphragm EPCs, a
thin circular diaphragm is attached to a fixed rigid annulus. As pressure is applied to the cell, the
diaphragm deflects and this deflection is sensed by one or more strain gauges, which then are
used to infer the applied pressure. For hydraulic EPCs, the operating principle involves two
plates that are welded together around the periphery and are filled with a fluid, e.g., hydraulic oil.
As external pressure is applied the fluid pressure inside the cell increases which is sensed by a
pressure transducer. Therefore, the pressure inside the cell infers the externally applied pressure.
Another less-common type of EPC is a null cell (Talesnick 2005, Talesnick et al. 2014,

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 82

Talesnick et al. 2021), which utilizes pneumatic back pressure that is applied inside the cell’s
cavity to ensure the diaphragm of the cell does not deflect (e.g., it remains in the “nulled”
position) under applied external stresses. The focus of the current study will be on deflection
based (e.g., diaphragm or hydraulic) EPCs, as they are more readily available compared to null-
based EPCs. One of the major challenges with EPCs that is well documented in the literature
(e.g., Sleig 1964, Weller and Kulhawy 1982) is the ability for an EPC to measure changes in
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

earth pressure in situ as if the EPC was not present (i.e., the presence of the sensor affects the
measurement results). Unfortunately, due to inherent stiffness differences between an EPC and
the surrounding soil medium, the pressures measured by an installed EPC do not match the
actual free stresses within the surrounding soil. Therefore, it is paramount that EPCs are
calibrated in a controlled laboratory environment that will resemble field conditions as closely as
possible (Brown 1977).
Typically, commercially available EPCs are delivered with a manufacturer-provided
calibration factor, which commonly has units of Electrical Measurement / Engineering Unit (e.g.,
mV/kPa, V/kPa). This calibration factor is typically determined by the manufacturer using a
laboratory calibration procedure that has an incompressible fluid (e.g., water, hydraulic oil, etc.)
as its medium to apply an external pressure to the EPC. The issue with utilizing an
incompressible fluid to calibrate an EPC is the fluid’s inability to account for soil-specific
phenomenon such as stress history, nonlinear load-unload hysteresis behavior, and the
development of shear stress (e.g., arching) along the perimeter of the cell (Labuz and Theroux
2005, Miller et al. 2007). To account for these factors, several studies have been conducted
where dry sand was utilized as the calibration medium prior to installing the EPC in the field.
Van Deusen et al. (1992) utilized a 1 cm thick-wall steel cylinder that was 35.6 cm in diameter
and 29.2 cm deep that was filled with dry sand as their loading chamber. The chamber was
loaded in a load frame where a 34.3 cm diameter load platen was connected to hydraulic actuator
in order to apply external pressure at the top surface of the loading chamber. For most of their
tests, the EPC was placed at depth of 14 cm within the loading chamber. For a given calibration
test where various load-unload cycles were applied to an EPC the authors noted the following:
1. The degree of hysteresis observed in the EPC’s response during loading and unloading
states decreases with increased load cycles.
2. The EPC’s maximum registration (e.g., voltage) response decreased with increased load
cycles for the same maximum applied external pressure.
Van Deusen et al. (1992) attributed the first factor to the inherent frictional behavior of soil
(especially dry sand). Similarly, Labuz and Theroux (2005) also experienced hysteresis behavior
during the unloading portion of their calibration tests, attributing this behavior to soil grains
interlocking and retaining a portion of the applied load during the unloading process. Van
Deusen et al. (1992) attributed the second factor above to the reduction of the soil’s void ratio
with increasing load cycle. This reduction in void ratio ultimately increases the shear strength
that is mobilized around the perimeter of the EPC, which causes the EPC’s diaphragm to deflect
less under the same applied pressure, thereby decreasing the EPC’s registration response. This
phenomenon is known as soil arching, and has been well-documented in geotechnical
engineering literature (e.g., Terzaghi 1943, McNulty 1965).
Similar behavior was observed in the study performed by Labuz and Theroux (2005), where
the authors performed laboratory calibration of EPCs in dry sand utilizing two different
experimental setups. The first apparatus consisted of placing an EPC inside the cavity of a steel
plate and placing a low-friction sleeve on top of the EPC’s sensing face, which either had a

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 83

height of 12.7 or 25.4 mm. The sleeve was filled with dry uniform sand and then a top cap was
placed over the EPC and sleeve. The column of dry sand above the EPC’s sensing face was
hermetically sealed through a rubber membrane. In order to apply external pressure to the soil
column which ultimately loaded the EPC’s sensing face, the top cap of the testing apparatus was
filled with hydraulic oil, which was pressurized during the loading sequence of each test. When
comparing the calibration factors determined from the aforementioned testing apparatus for
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

different testing conditions to the manufacturer’s calibration factor (e.g., fluid calibration), the
calibration factors of the EPC testing in dry sand conditions were consistently 20% lower relative
to the calibration factor provided by the manufacturer, indicating that soil arching had developed
when loading the EPC utilizing a dry sand column. To confirm this behavior the authors also
developed a full-scale testing apparatus utilizing a 55-gallon drum that had a 57.2 cm diameter
and was cut down to a height of 58.4 cm. Rubber bladders filled with fluid were placed at bottom
third and top third of the drum with dry sand filled in between them, with the EPC placed at the
center depth of the drum. The rubber bladders were utilized to ensure equal pressure was applied
at the top and bottom of the middle sand layer where the EPC was installed. Similar to Van
Deusen et al. (1992), the drum was loaded via a hydraulic actuator that applied load to a series of
stacked steel plates, with the largest plate having a diameter of 53.3 cm. The authors indicated
that the applied pressure from the actuator was approximately equal to the force applied by the
actuator divided by the diameter of the drum. This was validated by the pressure readings within
the rubber bladders, which the authors indicated were within 2% of each other. Similar to the
results found utilizing the first testing apparatus, the second full-scale testing apparatus revealed
that the calibration factors were consistently lower than the manufacturer’s calibration factor.
More specifically, when utilizing the manufacturer’s calibration factor to determine the pressure
sensed by the EPC, the pressure sensed by the EPC was consistently 20% less than the applied
stress of the system. Therefore, the manufacturer’s calibration factor (e.g., fluid calibration) did
not account for soil arching, resulting in an incorrect determination of pressure.
More recently, Miller et al. (2007) performed laboratory calibrations of EPCs in dry sand
utilizing the full-scale testing apparatus developed in Labuz and Theroux (2005). In their study,
the authors calibrated two EPCs with different pressure ranges of 100 kPa and 400 kPa,
respectively. Upon completion of the laboratory calibration of both EPCs, the authors calculated
a registration ratio, which was defined as the ratio between the manufacturer’s sensitivity ratio
and the dry sand’s sensitivity ratio. The registration ratios were 1.153 and 1.196 for the 100 kPa
and 400 kPa range EPCs, respectively, indicating that the measurements made by the EPCs when
placed in dry sand under-registered pressure values by roughly 15-20% relative to values that
would have been measured if the manufacturer’s calibration was used. This observation is
consistent with the aforementioned studies presented in this paper. The one limitation with the
study conducted by Miller et al. (2007) is that the authors utilized a total of 14 EPCs for other
studies (e.g., Mooney and Miller 2009, Rinehart and Mooney 2009), with pressure ranges of 100
kPa, 250 kPa, 400 kPa, and 1 MPa. However, the authors did not calibrate all of their EPCs in
dry sand, indicating that for a given pressure range, EPCs have shown extremely uniform results
when calibrated in dry sand. Based off of this information and utilizing linear interpolation the
authors estimated the registration ratios for EPCs having rated pressure values of 250 kPa and 1
MPa to be 1.175 and 1.291, respectively.
Ultimately, the current literature has indicated that it is imperative to calibrate EPCs in a
controlled laboratory environment with the EPC subjected to pressures applied by a column of
dry sand in order to accurately account for soil-specific phenomenon such as arching, stress

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 84

history, and nonlinear load-unload hysteresis behavior. The one limitation with the previous
studies conducted is their use of specially fabricated testing apparatuses in order to calibrate
EPCs in a laboratory environment, as it can be difficult to replicate the exact testing set up in
order for other researchers to independently calibrate their EPCs. Consequently, for the current
study, a simplified testing apparatus/method was developed to calibrate EPCs in dry sand. Three
1-MPa-range semiconductor EPCs were calibrated in dry sand at various sand column heights
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ranging from 5 cm to 30 cm, which were each subjected to ten loading cycles at each column
height. The goal of this study was to examine the influence of sand column height and load
cycles on the EPC’s associated sensitivity ratio (e.g., the ratio between the manufacturer’s
hydrostatic calibration factor and the measured soil calibration factor), while also providing
interested readers with a guide on how to calibrate an EPC utilizing readily available equipment
that is commonly found in most geotechnical testing laboratories.

EXPERIMENTAL SETUP

Earth Pressure Cells and Testing Material

Three Geokon 3510 series EPCs (Geokon 2019) were calibrated in this study. These EPCs
have a 100 mm diameter sensing face, a thickness of 10 mm, and a measurement range of 1
MPa. In order to sense changes in pressure, these EPCs are filled with an incompressible de-
aired hydraulic fluid. Changes in fluid pressure inside the EPC are sensed by a pressure
transducer that is semiconductor in nature, which converts this change in pressure to an electrical
signal. These EPCs were selected based off the recommended criteria outlined in Weiler and
Kulhawy (1982) along with their successful implementation to monitor dynamic changes of earth
pressures in situ by other researchers (e.g., Rinehart and Mooney 2009). The dry sand used to
calibrate each of the EPCs was a poorly graded silica sand, which is typically used with sand
cone testing for earthwork compaction QA/QC purposes (ASTM D1556-07). The silica sand
utilized in this study had a mean grain size (D50) of 0.60 mm resulting in a diameter ratio
between the EPC and the mean grain size (e.g., dEPC/D50) to be over 160, which was greater than
the ratio of 10 recommended by Weiler and Kulhawy (1982).

Testing Apparatus

As previously mentioned, the goal of this study was to develop a simplified testing procedure
in order to independently calibrate EPCs in dry sand without the need for custom-fabricated
testing equipment. To accomplish this, a 22.2 kN triaxial load frame configured with a 22.2 kN
load cell was utilized as the loading apparatus. In order to subject the EPCs to external pressures
applied through a soil column, a 152 mm inner diameter PVC pipe was cut to create various
column segments having heights ranging from 10 cm to 40 cm; these column segments were
then filled with sand and utilized to load the EPCs. For each specified sand column height, the
total height of the PVC column was an additional 5 cm, to account for the total thickness of the
EPC (sensing face and rigid back plate) and to allow a portion of the loading plate to stick out
from the column. Therefore, for a 10 cm sand column height, the PVC column utilized was 15
cm in height. In order to seal the sensing face of the EPC inside each PVC column, a vertical slit
was cut along the PVC column to allow the EPC sensing face to fit neatly inside the column.
Two latex membranes were glued together and were then glued along the circumference of each

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 85

PVC column. Once the EPC sensing face was secure inside the PVC column, the vertical slit and
outside base of the column was sealed utilizing electrical tape. In order to axially load each EPC
subjected to a column of sand, two steel plates with diameters of 11.4 cm and 15 cm were used
to transfer the load from the triaxial frame to the EPC. For this configuration, the applied
pressure during testing was assumed to be equivalent to the load applied by load frame divided
by the area of the 15 cm diameter loading plate, similar to the approach utilized by Van Duesen
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

et al. (1992) and Labuz and Theroux (2005). Figure 1 shows the materials that were used to
construct the PVC columns along with a typical EPC being loaded using the load frame and sand
column apparatus.

Figure 1. EPC calibration set up: a) equipment utilized, and b) EPC loaded in load frame.

Testing Procedure

Similar to the work done by previous researchers (e.g., Labuz and Theroux 2005), the EPCs
were loaded statically in 200 kPa increments up to 900 kPa (e.g., 200, 400, 600, 800, and 900
kPa). The loading of each EPC was controlled through the load frame’s software, which is
typically employed to conduct compressive triaxial tests on soil specimens. The response of the
EPC (e.g., voltage output) during loading was recorded by connecting the EPC to a Dataq DI-
2108-P datalogger, which scanned the response of the EPC at a frequency of 10 Hz. For a given
loading increment, the EPC’s response was recorded manually as it stabilized, once the system
had reached equilibrium. In order to determine the calibration factor for a given load cycle, a
simple linear regression was conducted between the EPC’s voltage response and the associated
applied pressure from the loading apparatus. The slope from this regression analysis was taken to
be the calibration factor, which was reported in units of kPa/volt in order to be consistent with
the calibration factor provided by the manufacturer. Figure 2 illustrates the nominal loading
regime for a given load cycle along with the associated linear regression that was used to
determine the calibration factor.
In order to account for soil-specific phenomenon such as arching and stress history, each
EPC was tested at various sand column heights consisting of 5, 10, 15, 20, and 30 cm. For a
given column height, 10 load cycles were applied to each EPC. The loading regime illustrated in
Figure 2a was repeated ten times for each calibration test at a given sand column height. This
number of load cycles was utilized to account for the fact that during the experimental set up for
each test, dry sand was poured in the PVC columns with in an initial density of approximately

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 86

1.66 g/cm3 resulting in the dry sand being in a relatively loose state prior to testing. Previous
research conducted by Labuz and Theroux (2005) subjected EPCs to five to six load cycles when
calibrating EPCs in similar material at an initial density of approximately 1.64 g/cm 3. Therefore,
the authors of this study deemed that 10 load cycles were a sufficient number of cycles for each
series of calibration tests.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. EPC calibration: a) loading regime, and b) associated linear regression analysis.

Limitation of Proposed Testing Methodology

One of the potential disadvantages of the proposed testing methodology is the proximity of
rigid boundaries to the EPC during the testing process. This is due to the fact that the diameter of
the EPCs tested within this study are 10 cm while the diameter of the PVC columns were 15 cm,
resulting in the perimeter of the EPCs only having a nominal 2.5 cm clearance away from the
inside wall of the PVC column. Given that EPCs are installed in situ to monitor earthen
pressures, EPCs should ideally be calibrated in a laboratory environment that resembles in situ
conditions, e.g., free-field conditions. However, given that the goal of this study was to develop a
simplified calibration procedure utilizing readily available equipment, a 15 cm diameter PVC
column was chosen as it was small enough to fit inside the triaxial load frame. Though larger
diameter PVC columns can be constructed (e.g., 30 cm diameter) this would require a much
larger triaxial load frame (e.g., 179.2 kN), which are not as readily available relative to smaller
triaxial load frames, thus ultimately defeating the purpose at attempting to develop a simplified
testing methodology that could allow other researchers to independently calibrate EPCs. It
should also be mentioned that the proposed testing apparatus is similar to the first testing
apparatus developed by Labuz and Theroux (2005). In this study, the authors utilized a sand
column only as wide as the diameter of the EPC to calibrate their EPC in dry sand. As mentioned
earlier in this paper, the results from Labuz and Theroux’s (2005) simplified experimental set up
were consistent with the results from calibration tests conducted using Labuz and Theroux’s
(2005) full-scale EPC testing apparatus. Therefore, as long as the entire face (e.g., sensing face
and inactive rim) of the EPC is subjected to loading via a sand column for calibration purposes,
the rigid boundaries imposed by the experimental test set up may not significantly influence the
calibration results. As this conclusion is beyond the scope of the current study, further
investigations are warranted to validate this assumption.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 87

RESULTS

Following the approach utilized by previous researchers (e.g., Van Deusen 1992, Miller et al.
2007), a calibration ratio was calculated for each load cycle utilizing Equation 1:

SenManufacturer
CR =
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(1)
Sensoil

where Sensoil is the sensitivity of the EPC in mV/kPa that was determined for each load cycle as
illustrated in Figure 2, SenManufacturer is the sensitivity of the EPC in mV/kPa provided by the
manufacturer that corresponds to the EPC’s response when pressurized by an incompressible
fluid, and CR is the resulting calibration ratio. For the EPCs calibrated in this study, the
manufacturer specified the gauge factor to range between 199.9 to 200.2 kPa/volt for each EPC,
resulting in a nominal EPC sensitivity of 5 mV/kPa. Figure 3 shows the variation in CR with
each load cycle applied to each of the EPCs calibrated in this study, for a sand column height of
10 cm.

Figure 3. Calibration results for EPCs with a 10 cm sand column height.

As shown in Figure 3, each of the EPCs that was calibrated experienced a reduction in CR
with each successive load cycle, with the CR for each EPC tending to stabilize fairly well after
10 load cycles. This phenomenon of a reduction in CR with increase in load cycle can be
attributed to the dry sand’s reduction in void ratio (and increase in relative density) with each
successive load cycle, which results in an increase in modulus of elasticity (Patel and Joshi
2021). This increase in modulus of elasticity has been shown to cause a reduction in registration
ratio e.g., CR (Weiler and Kulhawy 1982). When examining the results more closely, EPC 1 had
an initial CR of 1.163 and a final CR of 1.093 after 10 load cycles resulting in a 6.0% reduction
in CR. EPC 2 and 3 showed similar behavior with initial CR values of 1.098 and 1.104
respectively, and final CR values of 1.08 and 1.072 after 10 load cycles, resulting in a 1.64% and
2.9% reduction in CR, respectively. To provide context, the CR values determined by Miller et
al. (2007) utilizing similar EPCs were 1.153 and 1.196 after two load cycles when their EPCs

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 88

were subjected to loading with a sand column height of approximately 10 cm (e.g., Figure 8 in
Labuz and Theroux 2005). The CR values for each EPC in this study at a sand column height of
10 cm after two load cycles were 1.143, 1.111, and 1.106, indicating that the results from this
study are consistent with the results presented by other researchers (e.g., Van Deusen 1992,
Labuz and Theroux 2005). When comparing results, it should be noted that Miller et al. (2007)
only calibrated EPCs with ranges of 250 kPa and 400 kPa. However, based on their CR results,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

they extrapolated their CR value for a 1 MPa range EPC and determined it to be 1.291 (e.g., CR
is influenced by the pressure range of an EPC). The results from this study indicated that the CR
is not influenced by the pressure range of EPC, as the EPCs calibrated in this study had a
pressure range of 1 MPa and showed similar behavior to the 100 kPa and 400 kPa pressure range
EPCs that were calibrated in Miller et al. (2007) under similar testing conditions.
Another important aspect of the results shown in Figure 3 is that EPC 1 showed slightly
different behavior relative to EPC 2 and EPC 3. More specifically, EPC 1’s CR appeared to be
more sensitive to the volume of load cycles, as its CR saw a greater decrease with load cycles
relative to EPC 2 and EPC 3. This is an important finding as Miller et al. (2007) indicated that
other researchers have shown extreme uniformity between EPCs of the same pressure range
when calibrating their EPCs in dry sand. The results in Figure 3 suggest that in some instances
(e.g., EPC 2 and EPC 3) this assumption holds true, while in other instances (e.g., EPC 1 relative
to EPC 2 and 3) this may not be a valid assumption. It is important to note, however, that CR
values for EPC 1 did not drastically differ compared to the CR values of EPC 2 and EPC 3. For a
sand column height of 10 cm, the CR values for a given load cycle for EPC 1 relative to EPC 2
and EPC 3 were on average 2.73% higher with the highest and lowest difference being 5.92%
and 1.2%. Ultimately, the results in Figure 3 suggest that it is advantageous to calibrate each
EPC individually if multiple EPCs are to be utilized for in situ applications, as though the trends
between EPCs are similar, some EPCs might exhibit slightly different behavior relative to other
EPCs having the same pressure range.
Figure 4 shows a summary of the results for this study, indicating the variation of CR with
sand column height and load cycle. Variation in sand column height was investigated to
determine if a critical sand column height existed where no further arching would develop with
additional increases in column height, which would manifest itself as no further changes in CR.
Previous researchers (Labuz and Theroux 2005) have indicated that this “critical height” should
be twice the diameter of the sensing face of the EPC, which is why sand column heights of 20
cm and 30 cm were investigated in the current study. Figure 4 clearly illustrates that CR
continually increases with increasing column height, indicating that a critical sand column height
was not achieved using this experimental set up. Figure 4 also illustrates that with increasing
sand column height, the variation in CR with load cycle becomes more pronounced. The authors
attribute this behavior to the increase in the volume of voids that occurs as the sand column
height increases. Given that the dry sand is prepared in the same manner for each column height,
resulting in an initial density of 1.66 g/cm3, an increase in sand column height results in an
increase in the total void space within the sand column. This allows for a greater change in void
ratio during each successive load cycle for sand columns of greater height, which results in a
greater change in CR with load cycles. More specifically, sand column heights of 10 cm and 15
cm, on average, experienced a 3.5% and 2.5% reduction in CR between the first and tenth load
cycle respectively. On the other hand, sand column heights of 20 cm and 30 cm experienced a
34.3% and 31.4% reduction in CR between the first and tenth load cycle respectively. It is also
interesting to point out that a sand column height of 5 cm experienced a CR of less than 1 for all

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 89

three of the EPCs. This finding is consistent with the findings from Van Deusen et al. (1992),
who determined a CR of 0.70 when calibrating an EPC at a column height of 7.5 cm. Van
Deusen et al. (1992) ultimately attributed this behavior to the EPCs relative distance to the
loading plate utilized in their study, indicating that geotechnical instruments such as EPCs should
maintain a reasonable distance away from rigid boundaries, as this may affect the response of the
EPC.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Ultimately, the results from this study (Figure 4) indicated that sand column heights of 10 cm
and 15 cm produced the most reliable calibration results, which is a finding that is consistent
with the current literature. These findings also have practical significance as Theroux et al.
(2001) have indicated that the installation of an EPC in situ should mimic the conditions in
which it was calibrated in the laboratory. Installing an EPC in situ with a sand column height of
10 cm or 15 cm is more practical in the field relative to a sand column height of 20 cm or 30 cm.
Therefore, it is recommended that these types of EPCs be calibrated with sand column heights of
either 10 cm or 15 cm if they are to be installed in the field for later use.

Figure 4. Summary of all calibration results measured in this study.

Assessment of Repeatability

Initial results from this study indicated that the CR for these EPCs calibrated with a 10 cm
sand column height ranged from 1.16 to 1.07. However, these results were only determined from
one series of tests for each EPC. In order to gain a better understanding of the repeatability of the
proposed experimental testing program, EPC 3 was subjected to five additional testing sequences
as outlined in the aforementioned Testing Procedure section. A 10 cm column height was
ultimately chosen for this repeatability testing, as initial results revealed that this column height
provided relatively consistent results between EPCs, and from a practical point of view a 10 cm
column height would be manageable to replicate in the field, more so than taller sand column
heights. Figure 5a provides a histogram of the CR results obtained from the repeatability testing

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 90

for all ten loading cycles for a given testing sequence. Similar to the initial results, the average
CR value for these series of tests is 1.17, which is consistent with previous literature for an EPC
tested at a similar sand column height. The coefficient of variation from this series of testing is
relatively low at 4.6%, and the CR values ranged from 1.05 to 1.31 with the middle 95% of the
CR values ranging from 1.06 to 1.28 based off the 2.5% and 97.5% percentile, respectively.
Given that CR experiences greater changes during the initial load cycles of a given test sequence,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5b provides a histogram of the CR results obtained from the repeatability testing for the
last five load cycles of each sequence of tests. As shown, CR saw a reduction in mean, range, and
middle 95% values with values of 1.14, 1.05 to 1.20, and 1.06 to 1.20, respectively. Ultimately,
the results from this repeatability study reveal one of the potential disadvantages with this
proposed simplified experimental procedure. Though the results outlined in Figure 3 appear to be
promising, further results presented in Figure 5 provide a better illustration of the potential
variability that other researchers may encounter. However, the CR values determined from this
portion of the study were still consisted with the CR values determined by previous researchers
(Labuz and Theroux 2005, Miller et al. 2007).

Figure 5. Results from repeatability testing for EPC3: a) all load cycles per testing
sequence, and b) last five load cycles per testing sequence.

CONCLUSIONS

Previous research has indicated that EPCs tend to under register earth pressures when
utilizing the manufacturer’s calibration factor, which is determined by calibrating the EPC in an
incompressible fluid. When an installed EPC is subjected to loading in place, the shear strength
of the soil around the perimeter of the cell is mobilized as the sensing face of the EPC deforms
relative to the stiff outer rim of the EPC, resulting in a stress arching effect. This behavior
necessitates that EPCs be independently calibrated in dry sand if this soil-specific phenomenon is
to be reasonably accounted for. Previous researchers have performed soil calibrations of EPCs
using specially fabricated testing apparatuses and have shown that the calibration factor
determined in dry sand differs by 10-20% relative to the manufacturer’s calibration factor. The
one limitation with the previous studies that have been conducted is the use of custom-fabricated
testing apparatuses for calibrating EPCs in a laboratory environment, which ultimately makes it
difficult for other researchers to replicate the exact testing setup. In the current study, a very

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 91

simple testing apparatus was developed in order to calibrate three 1-MPa-range EPCs, by
utilizing PVC columns cut at various heights to house the sand column above the EPC sensing
face. For this setup, a 5-kN load frame was used to load the soil and pressurize the EPC. The
following conclusions were drawn from calibration tests conducted using this simplified
calibration approach:
1. Under the same testing conditions as previous research (e.g., the same sand column
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

height), the ratios between the soil calibration factor and the manufacturer’s calibration
factor within this study were consistent with the ratios determined from previous
research. More specifically, the ratios determined from this study when pressurizing an
EPC at a sand column height of 10 cm ranged from 1.07 to 1.16, while previous research
indicated ratios that ranged from 1.15 to 1.19.
2. The CR value for a given sand column height tended to fluctuate between the three EPCs
that were calibrated in this study, indicating that the assumption that EPCs of the same
pressure range show extreme uniformity under the same loading conditions may not
always be valid. Within this study, at a sand column height of 10 cm, EPC 1 had CR
values that ranged from 1.16 to 1.09 under 10 load cycles, while EPCs 2 and 3 had CR
values that ranged from 1.09 to 1.08 and from 1.10 to 1.07, respectively. Although it can
be argued that these are not significant differences, the results from this study suggest
that it is advantageous to calibrate each EPC independently for more accurate field
results.
3. When investigating the influence of sand column height on CR for sand columns up to 30
cm in height, an increase in sand column height resulted in an increase in CR, and also
revealed a greater change/reduction in CR for each successive load cycle for a given test.
More specifically, when the three EPCs were calibrated under a 10 cm sand column
height, the associated CRs saw on average a 3.5% reduction after 10 load cycles, while
under a 20 cm sand column height the associated CRs saw on average a 34% reduction
under the same number of load cycles. These results have practical significance, as EPCs
instrumented in situ should be instrumented in a manner that best replicates the
conditions in which they were calibrated in a laboratory. Therefore, instrumenting an
EPC with a sand column height of 10 cm in the field is much easier to achieve as oppose
to an EPC with a sand column height of 20 cm.
4. A repeatability study was performed in order to assess the reliability of the proposed
testing procedure and the variability that may be encountered when calibrating a single
EPC in dry sand. Results from five duplicate testing sequences for a sand column height
of 10 cm revealed an average CR value of 1.17 and a coefficient of variation of 4.6%.
When only considering the last five load cycles from each testing sequence, both the
average CR value and coefficient of variation between results reduce to 1.14 and 3.7%,
respectively. Overall, the results from the repeatability study revealed the potential
variation that can be encountered when conducting this proposed experimental setup in
order to calibrate an EPC in dry sand. The CR values obtained from this portion of the
study are consistent with the initial results presented from this study, and the results
presented in the literature.
Overall, the results from the current study were consistent with results presented by previous
researchers for EPCs calibrated at the same sand column height, indicating that this simplified
testing procedure may be appropriate for relatively rapid, inexpensive, and independent
calibration of EPCs.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 92

ACKNOWLEDGEMENTS

This material is based upon work supported by the Delaware Department of Transportation
under DelDOT Project Number T201966002, Task 1891-28.

REFERENCES
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ASTM. ASTM D1556-07. (2007). Standard Test Method for Density and Unit Weight of Soil in
Pl. by the Sand-Cone Method. Annual Book of ASTM Standards, Vol. 04.08, ASTM
International, West Conshohocken, PA.
Brown, S. F. (1977). “State-of-the-Art Report on Field Instrumentation for Pavement
Experiments.” Transportation Research Record, 640, 13-28.
Brown, S. F., and Brodrick, B. V. (1981). “Instrumentation for the Nottingham Pavement Test
Facility.” Transportation Research Record, 810, 73-79.
Dawson, A. R., and Little, P. H. (1997). “Measurement of Stress and Strain in an Unsurfaced
Haul Road at a Soft Clay Site in Scotland.” Transportation Research Record, 1596, 15-22.
Geokon. (2019). “Model 3500 Series Earth Pressure Cells”. Instruction Manual, Revised
04/30/2019, Geokon, Lebanon NH.
Labuz, J. F., and Theroux, B. (2005). “Laboratory Calibration of Earth Pressure Cells,”
Geotechnical Testing Journal, Vol. 28, No. 2, 1-9.
McNulty, J. W. (1965). “An Experimental Study of Arching in Sand.” Technical Report No. 1-
674, U.S. Army Engineer Waterways Experiment Station, Vicksburg, MS.
Miller, P. K., Rinehart, R. V., and Mooney, M. A. (2007). “Measurement of Soil Stress and
Strain using In-ground Instrumentation.” Proceeding of GeoDenver 2007: New Peaks in
Geotechnics, Geotechnical Special Publication No. 160, February 18-21, 2007, Denver, CO,
ASCE, Reston, VA, 1-10.
Patel, A. V., and Joshi, N. H. (2021). “Effect of Relative Density on Elastic Properties of Sand.”
Proceedings of the Indian Geotechnical Conference 2019. Lecture Notes in Civil
Engineering, Vol 133. Springer, Singapore.
Rinehart, R. V., and Mooney, M. A. (2009). “Measurement of Roller Compactor Induced
Triaxial Soil Stresses and Strains.” Geotechnical Testing Journal, 32(4), 347-357.
Selig, E. T. (1964). “A Review of Stress and Strain Measurements in Soil.” Proceedings of the
Symposium on Soil-Structure Interaction, University of Arizona, 172-186.
Talesnick, M. (2005). “Measuring Soil Contact Pressure on a Solid Boundary and Quantifying
Soil Arching.” Geotechnical Testing Journal, 28(2), 171-179.
Talesnick, M. L., Avraham, R., and Ringel, M. (2014). “Measurement of Contact Soil Pressure
in Physical Modelling of Soil-Structure Interaction.” International Journal of Physical
Modelling in Geotechnics, 14(1), 3-12.
Talesnick, M., Ringel, M., and Rollins, K. (2021). Development of a Hybrid Soil Pressure
Sensor and its Application to Soil Compaction.” Canadian Geotechnical Journal, 58(6), 811-
822.
Terzaghi, K. (1943). Theoretical Soil Mechanics, Wiley, New York.
Theroux, B., Labuz, J. F., and Dai, S. (2001). “Field Installation of an Earth Pressure Cell.”
Transportation Research Record: Journal of the Transportation Research Board, 1772(1),
12-19.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 93

Van Deusen, D. A., Newcomb, D. E., and Labuz, J. F. (1992). “A Review of Instrumentation
Technology for the Minnesota Road Research Project.” Final Report – Minnesota
Department of Transportation Report No. FHWA/MN/RC – 92/10.
Vennapusa, P. K. R., White, D. J., Siekmeier, J., and Embacher, R. A. (2012). “In situ
Mechanistic Characterisations of Granular Pavement Foundation Layers.” International
Journal of Pavement Engineering, 13(1), 52-67.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Weller, W. A., Jr., and Kulhawy, F. H. (1982). “Factors Affecting Stress Cell Measurements in
Soil.” Journal of the Geotechnical Engineering, 108(12), ASCE, 1529-1548.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 94

HVSR Measurements to Investigate Sinkholes and Treatment Efforts along a Roadway

Pourya Alidoust1; Siavash Mahvelati, Ph.D.2; Joseph T. Coe, Ph.D.3; Atsuhiro Muto, Ph.D.4;
Sarah McInnes, P.E.5; Mia Painter, P.G.6; and Katherine Kubiak7
1
Dept. of Civil and Environmental Engineering, Temple Univ., Philadelphia, PA.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Email: pouryaalidoust@temple.edu
2
Vibra-Tech Engineers, Inc., Hazleton, PA. Email: siavashm@vibratechinc.com
3
Dept. of Civil and Environmental Engineering, Temple Univ., Philadelphia, PA.
Email: joseph.coe@temple.edu
4
Dept. of Earth and Environmental Science, Temple Univ., Philadelphia, PA.
Email: amuto@temple.edu
5
Pennsylvania Dept. of Transportation, King of Prussia, PA. Email: smcinnes@pa.gov
6
Schnabel Engineering, Chadds Ford, PA. Email: mpainter@schnabel-eng.com
7
Dept. of Civil and Environmental Engineering, Temple Univ., Philadelphia, PA.
Email: katherine.kubiak@temple.edu

ABSTRACT

The evaluation of sinkholes continues to present challenges when managing highway


infrastructure given current geotechnical subsurface investigation techniques. Typical drilling
efforts combined with standard penetration testing (SPT) and cone penetration testing (CPT)
provide useful information, but only at a limited number of points throughout the site. Such
efforts are also not routinely utilized after the implementation of sinkhole treatment efforts as a
means of quality control due to logistical challenges and costs. The horizontal-to-vertical spectral
ratio (HVSR) method, which records ambient seismic noise using a single, broad-band three-
component seismometer, can be rapidly deployed throughout a site and provides information
regarding the stiffness of the subsurface materials. Repeated HVSR measurements at the same
locations can potentially identify problematic areas present after the implementation of sinkhole
treatment efforts. In this paper, HVSR measurements were collected along a roadway where
known sinkholes were present. The HVSR testing locations were collocated with boreholes to
allow for a comparison with the inferred subsurface profile from geotechnical investigations. The
roadway was then subjected to an extensive grouting program to treat the active sinkholes.
HVSR measurements were subsequently repeated at the same pre-grouting locations to
investigate changes in the subsurface stiffness. The results exhibited an increase in HVSR
amplitude toward lower frequencies when comparing the pre- and post-grouting ambient noise
measurements, possibly indicating a stiffening of the soil post-grouting. Despite rejection of time
windows that exhibited evidence of near-field anthropogenic waveforms, there was still
significant variance in the results attributed to poor coupling of the seismometer (i.e., pavement
surface and gravelly fill on the roadway shoulder), adverse weather conditions (i.e., wind and
rain), and the complexity of the site conditions (i.e., the extent and spatial variability with which
grout flowed into voids and interaction of grout with nearby buried utilities). Use of HVSR as a
quality control tool for sinkhole treatment efforts should be further investigated with
consideration of the complexities that can affect such variance in the results.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 95

INTRODUCTION

The internal erosion and dissolution of soluble evaporate or carbonate rocks leads to the
formation of funnel-shaped depressions known as sinkholes. Sinkholes are the most diagnostic
representation of karst terrain, which varies significantly in severity and poses a substantial
hazard to the stability of infrastructure. The financial impact of catastrophic land subsidence due
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

to sinkholes is significant. For example, the average cost of karst-related damages in the United
States has been conservatively estimated in the past to be at least $300,000,000 per year (Weary
2015). In the United States, 20 states are under significant danger of karstic activity, and more
than 17,000 square miles in 45 states have been directly affected by karst-related subsidence
(Galloway et al. 1999). The extent and variety of karst terrain across the United States
necessitate detailed karst hazard assessment with a special focus on mitigation and monitoring
plans.
Mitigation of karst usually involves stabilizing the underlying soil against erosion with
different ground improvement techniques such as grouting or compaction. Due to the distinctive
geological and hydrological characteristics of sinkholes, remedial measures to repair the existing
anomalies and prevent their potential reactivation must be uniquely tailored (Zeng and Zhou
2019). Poorly managed treatment projects can exacerbate the existing conditions, which can lead
to the development of additional sinkholes and groundwater contamination (Zhou and Beck
2008). Thus, evaluating the effectiveness of technologies used to remedy a sinkhole is of great
importance. Boreholes can be used to detect sinkholes and to evaluate the efficiency of an
implemented mitigation method. For example, very low SPT blow counts or loss of drilling fluid
circulation are typically indicative of karst-related void features. However, direct sampling
methods only provide information at select spots that cannot be reliably generalized to larger
areas. Alternative methods of investigation involve the precise measurement of subsidence rates
using landslide monitoring instruments such as Interferometric Synthetic Aperture Radar
(InSAR) or high precision leveling (Buchignani et al. 2008, Jones and Blom 2014, Gutiérrez et
al. 2019). However, these techniques are not prevalent in practice due to the sudden and
unpredictable nature of sinkholes (Guerrero et al. 2008) and logistical challenges and associated
costs related to implementation.
As a potential solution, the Horizontal-to-Vertical Spectral Ratio (HVSR) method can be
rapidly deployed at a site and provide information regarding the stiffness of the underlying
geologic materials. In the HVSR method, ambient seismic noise is recorded using a single broad-
band, three-component seismometer. This passive seismic method has continually evolved over
the past six decades (Nogoshi 1971, Nakamura 1989) and has been increasingly used to study
ground motion amplification as attempts have been made at its standardization (e.g., SESAME
2004). The ambient noise is typically recorded for a period of at least 30 minutes. For
processing, the entire recording is broken into smaller windows of 10-40 seconds. Windows
containing clear active or transient seismic signals are removed from the analysis and only those
with ambient vibrations are kept. The ratio between the Fourier amplitude spectra of the
horizontal (H) to vertical (V) components of the waveforms is then computed (i.e., H/V curve).
This process is repeated for each window, which results in a set of H/V curves that are
statistically examined to generate a representative H/V curve and uncertainty measurement. The
predominant frequency of the H/V curve (fo,HVSR) occurs at the largest H/V amplitude. The fo,HVSR
is interpreted to be the natural (resonant) frequency of the site (f0) (Molnar et al. 2022). HVSR
has shown promising results in providing consistent information related to site response effects,

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 96

especially when the subsurface contains strata with high impedance contrast (Mucciarelli 1998,
Bonnefoy-Claudet et al. 2006).
This study explores the applicability of the HVSR technique as a potential sinkhole detection
and treatment monitoring method. Ambient noise measurements were collected along a roadway
where known sinkholes were present. The locations of the HVSR measurement were collocated
with boreholes to allow for a comparison with the inferred subsurface profile from geotechnical
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

investigations. The roadway was then subjected to an extensive grouting program to treat the
active sinkholes. HVSR measurements were subsequently repeated at the same pre-grouting
locations to investigate for changes in the subsurface stiffness.

SITE DESCRIPTION

Large areas of central and eastern Pennsylvania are underlain by carbonate bedrock
(Kochanov 1999). The test site for this study exhibited reoccurring surface subsidence caused by
multiple sinkholes along State Route 3015 (Chemical Road) in Montgomery County,
Pennsylvania (Figure 1). Based on the geological data, the site is underlain by a dolomite
formation that is highly prone to karstic activity and requires special attention to treat sinkholes
during/after construction (Barnes 2008). Karst-related activity at this site led to multiple asphalt
repairs and treatment efforts in recent years. Several geophysical and geotechnical
investigations were conducted to investigate the subsurface and evaluate the sinkhole activity.
This included test borings that were drilled to aid in evaluating treatment strategies. The
locations of a subset of these borings and the corresponding interpreted subsurface profile are
provided in Figure 1. The SPT blow counts suggest a stiffness (i.e., shear wave velocity, VS)
reversal in the upper gravel/sand layers above the underlying dolomite rock. This stiffness
reversal was particularly prominent in boreholes R-11 and R-12 and the very low blow counts
likely result from the nearby active sinkholes. Given the subsurface investigation results, a
limited mobility grouting (LMG) program was selected to stabilize the ground and address the
sinkhole activity.

DATA COLLECTION AND PROCESSING

The ambient noise measurements in this study were performed using a three-component
Tromino® seismometer, a common tool employed in many HVSR studies (Giacomo and
Gallipoli 2005, Gosar 2007, Bignardi et al. 2016). The seismometer was placed immediately
adjacent to the location of previous boreholes when deployed on the roadway asphalt (Figure 1).
The seismometer was also deployed on soil along the roadway shoulder to acquire an additional
set of recordings in case the asphalt negatively affected the H/V curves. To obtain good
coupling, short spikes were used to couple the seismometer to the asphalt, while longer ones
were attached for deployment on soil. Additionally, the seismometer was covered with a bucket
to reduce the effect of wind, and placed inside a small hole in the topsoil for recordings along the
shoulder. The recording duration was 30 minutes at a sampling frequency of 128 Hz for each
location. Following the completion of grouting, HVSR measurements were repeated at the same
locations to investigate changes in the subsurface stiffness based on changes to the H/V curves,
including fo,HVSR.
The open-source Python package hvsrpy was used to process the HVSR measurements
(Vantassel 2021). This package utilizes a lognormal distribution to characterize fo,HVSR and an

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 97

automatic frequency-domain window-rejection technique. This algorithm reduces variance and


improves data quality while not over-rejecting good windows in low-variance data sets (Cox et
al. 2020). Each recording was broken down into 40-second windows. Unsmoothed edges of the
truncated windows can cause artifacts during the computation of Fourier spectra (Dal Moro
2015). Therefore, each window was tapered with a cosine taper function (width = 0.1) and band-
pass filtered (0.1-80 Hz). Fourier amplitude curves were then computed for the vertical and
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

horizontal components. The Fourier amplitude spectra were smoothed as described in Konno and
Ohmachi (1998). The geometric mean method was used to combine the two horizontal spectra
into a representative horizontal spectrum. A set of H/V curves was produced for each 40-second
window by dividing the horizontal spectrum by vertical spectrum. H/V curves that exhibited
standard deviations greater than two were rejected. The rejection and re-calculation of statistics
was repeated for several iterations until convergence (i.e., no more windows rejected). Figure 2
shows an example of the data processing for one recording. More details on the automated
frequency-domain window-rejection algorithm and lognormal statistics are provided in Cox et al.
(2020).

Figure 1. Project site: (a) location of site, test borings, and sinkholes (outlined in blue); and
(b) interpreted subsurface profile [including blow counts and % Recovery (RQD)].

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 98

Notes:
LMcurve = lognormal median curve
LMcurve ± 1 STD = uncertainty in LMcurve
expressed with positive/negative standard
deviations
f0,i = peaks of accepted time windows
f0,mc = peak of the median curve
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

LMf0 ± 1 STD = uncertainty in LMf0 expressed


with positive/negative standard deviations

Figure 2. Sample recording processed using hvsrpy: (a) north-south (NS) component; (c)
east-west (EW) component; (e) vertical (VT) component; (b) H/V curves before rejection
algorithm; and (d) H/V curves after rejection algorithm.

RESULTS

Artificial materials such as asphalt or concrete are typically stiffer than underlying soil strata,
which creates a velocity inversion at the immediate near surface. In many HVSR studies
conducted in urban locations, deploying the seismometer on such materials can be impossible to
avoid. There is no clear consensus in the technical literature regarding the exact effects of
pavement on the subsequent characteristics of the recorded ambient vibrations. While some
studies reported little to minor changes in amplitude (Acerra et al. 2004, Chatelain et al. 2008),
others have suggested that stiff coupling layers such as pavement have a significant impact on
higher frequencies (Castellaro and Mulargia 2009). Consequently, the effects of seismometer
coupling were examined by comparing the H/V curves obtained at the same approximate
locations on the soil shoulder and roadway asphalt.
As noted in Figure 3a, only minor differences were observed in the processed H/V curves
derived from ambient noise measurements on asphalt and soil. The main exception was that
some of the post-grouting recordings along the soil shoulder exhibited very large H/V amplitudes
below approximately 2 – 3 Hz (Figure 3b). Variations in H/V amplitude below that frequency
range can largely be affected by weather conditions at the site (Gutenberg 1958) with higher
amplitudes attributed to adverse weather conditions (Castellaro 2016). Post-grouting testing at
the site did take place under increasingly adverse weather conditions as a storm approached the
site, particularly along the shoulder of the roadway. This included large wind gusts and periods
of sporadic rain showers. Figure 3 also shows that some recordings exhibited minor reductions in
H/V amplitude at higher frequencies for the asphalt recordings, similar to the observations

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 99

regarding stiff coupling layers in Castellaro and Mulargia (2009). All further discussion
regarding the HVSR results focus on the recordings obtained on asphalt since they were less
impacted by the post-grouting adverse weather conditions, and they were located closer to the
adjacent boreholes.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Figure 3. Comparison of processed H/V curves on the soil shoulder and asphalt roadway:
(a) adjacent to borehole R-15 (pre-grouting); and (b) adjacent to borehole R-11 (post-
grouting). Note: Dotted lines indicate the uncertainty in the H/V curve (LMcurve±1 STD).

Figure 4 presents the processed H/V curves and shows evidence of larger standard deviation
(LMcurve±1 STD) in the post-grouting measurements relative to the pre-grouting measurements.
This was the case despite rejection of time windows that exhibited evidence of near-field
anthropogenic waveforms. Possible sources for the additional uncertainty between post- and pre-
grouting measurements include differences in seismometer coupling, adverse weather conditions,
and complexity of subsurface conditions (e.g., the extent and spatial variability with which grout
flowed into voids, interaction of grout with nearby buried utilities, etc.). Such sources have been
noted in the literature to increase H/V uncertainty (e.g., Cox et al. 2020).
Even accounting for the larger scatter, Figure 4 shows a general increase in H/V amplitude at
lower frequencies after grouting, particularly next to the active sinkholes (boreholes R-11 and R-
12). Both pre- and post-grouting HVSR measurements were obtained on weekends during which
the site had similar levels of anthropic activity (e.g., no construction, similar nearby traffic, etc.).
The increase in H/V amplitude also occurs over a large span of frequencies and above the range
typically associated with adverse weather conditions (< 2 - 3 Hz). Consequently, the increase in
H/V amplitude can be reasonably attributed to physical changes in the subsurface despite the
increased uncertainty present in the post-grouting measurements. Stratigraphically, this increase
in H/V amplitude can be explained by a grouting-related velocity increase in the soft zones noted
in Figure 1. The resulting subsurface stiffness profile would more closely resemble a gradually
increasing pattern with depth, which has been shown in other studies to result in H/V amplitudes
larger than one over a broad range of frequencies (e.g., Castellaro and Mulargia 2009).
The increase in H/V amplitude occurs over a broad enough frequency that a distinct peak is
not readily visible in the results (Figure 4); thus, it hinders the extraction of a reliable fo,HVSR
estimate based on the six clarity criteria established in SESAME (2004). Prior to grouting, a
reliable estimate for fo,HVSR was possible based on SESAME (2004) for all test locations except

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 100

for borehole R-11 right next to one of the active sinkholes at the site. The lack of a distinct H/V
peak for post-grouting measurements can be attributed to the tendency of the grout to decrease
the severity of large impedance contrasts. The results near borehole R-15, however, were quite
different suggesting perhaps a localized anomaly or differences in grouting. The increase in H/V
amplitude was much more narrowly defined and a distinct peak could be reliably interpreted
based on SESAME (2004). The cause of this peak was interpreted as stratigraphic since
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

anthropogenic artifacts appear as sharp/narrow spikes in all three spectral components (NS, EW,
and VT), which was not observed in this recording. When compared to the pre-grouting fo,HVSR (≈
12 Hz), the post-grouting fo,HVSR shifted to a lower frequency (≈ 9 Hz), which implies VS
decreased throughout the subsurface (based on fo,HVSR = f0 = VS/4H and a consistent bedrock
depth, H). This is unlikely after the introduction of low mobility grout. Instead, it should be
noted that the microtremor wavefield is a complex function of many variables and fo,HVSR may
not always equal to f0, particularly when large stiffness contrasts are reduced in the subsurface
(Molnar et al. 2022). Interestingly, the shift in fo,HVSR near borehole R-15 could be explained
mathematically if the post-grouting operations altered the site response there such that fo,HVSR was
now directly related to the next higher resonant site frequency (i.e., f1 = 3VS/4H) instead of f0
(assuming VS and H did not drastically change). Additional modeling and/or joint inversion with
other geophysical data would be necessary to definitively ascribe the cause of the distinct H/V
peak present in the post-grouting measurements near borehole R-15.

(a) (b)

(c) (d)

Figure 4. Comparison of processed H/V curves for pre- and post-grouting conditions
(asphalt coupling) adjacent to (a) borehole R-10; (b) borehole R-11; (c) borehole R-12; and
(d) borehole R-15. Note: Dotted lines indicate the uncertainty in the H/V curve (LMcurve±1
STD).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 101

CONCLUSIONS

The results of this study demonstrate that treatment of sinkholes via grouting can appreciably
alter the site subsurface response to ambient seismic noise. Pre- and post-grouting HVSR
measurements exhibit differences that can be inferred to be stratigraphic in origin. The main
observations from this study include: (a) deployment of the seismometer on asphalt resulted in
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

only minor differences in the H/V curves when compared to measurements on soil; (b) post-
grouting H/V amplitudes were larger between approximately 1 Hz – 10 Hz, which suggested the
post-grouting subsurface conditions more closely resembled a gradually increasing stiffness with
depth; (c) post-grouting HVSR measurements exhibited larger scatter/uncertainty; and (d) the
introduction of grout into the subsurface complicates the nature of the underlying microtremor
wavefield, and the typical interpretation of fo,HVSR closely corresponding to f0 may no longer be
true. The last two points highlight the complexity of using HVSR as a quality control tool for
sinkhole treatment efforts. Additional modeling and field measurements are recommended to
further elucidate the effectiveness and robustness of HVSR when used in this application.

ACKNOWLEDGEMENTS

This work was sponsored by the Pennsylvania Department of Transportation and the U.S.
Department of Transportation, Federal Highway Administration. The contents of this report
reflect the views of the authors, who are responsible for the facts and the accuracy of the data
presented herein. The contents do not necessarily reflect the official views or policies of the
Federal Highway Administration, U.S. Department of Transportation, or the Commonwealth of
Pennsylvania at the time of publication. This manuscript does not constitute a standard,
specification, or regulation. The authors gratefully appreciate the support of Jeffrey L Sinski at
Exploration Instruments, LLC, Jeremy Brown at Schnabel Engineering, Inc., and Timothy
Homan at PennDOT.

REFERENCES

Barnes, J. (2008). Rocks and Minerals of Pennsylvania. 4th ed. Pennsylvania Geological Survey,
4th ser., Educational Series 1.
Bignardi, S., Mantovani, A., and Zeid, N. (2016). “OpenHVSR: imaging the subsurface 2D/3D
elastic properties through multiple HVSR modeling and inversion.” Comp. & Geosci., 93,
103–113.
Bonnefoy-Claudet, S., Cotton, F., and Bard, P. (2006). “The nature of noise wavefield and its
applications for site effects studies: A literature review.” Earth-Science Reviews, 79(3–4),
205–227.
Buchignani, V., Avanzi, G. D. A., Giannecchini, R., and Puccinelli, A. (2008). “Evaporite karst
and sinkholes: A synthesis on the case of Camaiore (Italy).” Environ. Geology, 53 (5), 1037–
1044.
Castellaro, S. (2016). “The complementarity of H/V and dispersion curves.” Geophysics, 81 (6),
T323–T338.
Castellaro, S., and Mulargia, F. (2009). “The effect of velocity inversions on H/V.” Pure & App.
Geophys., 166 (4), 567–592.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 102

Chatelain, J.-L., et al. (2008). “Evaluation of the influence of experimental conditions on H/V
results from ambient noise recordings.” Bull. of Earthquake Eng., 6 (1), 33–74.
Cox, B., Cheng, T., and Vantassel, J. (2020). “A statistical representation and frequency-domain
window-rejection algorithm for single-station HVSR measurements.” Geophys. J. Int., 221
(3), 2170–2183.
Dal Moro, G. (2015). “Joint analysis of Rayleigh-wave dispersion and HVSR of lunar seismic
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

data from the Apollo 14 and 16 sites.” Icarus, 254, 338–349.


Galloway, D., Jones, D., and Ingebritsen, S. (1999). Land subsidence in the United States.
Circular 1182, U.S. Geological Survey, Reston, Virginia, 15 pp.
Di Giacomo, D., and Gallipoli, M. (2005). “Analysis and Modeling of HVSR in the Presence of
a Velocity Inversion: The Case of Venosa, Italy.” Bull. Seis. Soc. America, 95 (6), 2364–
2372.
Gosar, A. (2007). “Microtremor HVSR study for assessing site effects in the Bovec basin
(NW Slovenia) related to 1998 Mw5. 6 and 2004 Mw5. 2 earthquakes.” Eng. Geology,
178–193.
Guerrero, J., Gutiérrez, F., Bonachea, J., and Lucha, P. (2008). “A sinkhole susceptibility
zonation based on paleokarst analysis along a stretch of the Madrid–Barcelona high-speed
railway built over gypsum-and salt.” Eng. Geology, 102 (1), 62–73.
Gutiérrez, F., Benito-Calvo, A., Carbonel, D., and Desir, G. (2019). “Review on sinkhole
monitoring and performance of remediation measures by high-precision leveling and
terrestrial laser scanner in the salt karst of the Ebro.” Eng. Geology, 248, 283–308.
Jones, C. and Blom, R. (2014). “Bayou Corne, Louisiana, sinkhole: Precursory deformation
measured by radar interferometry.” Geology, 42 (2), 111–124.
Kochanov, W. E. (1999). Sinkholes in Pennsylvania. Pennsylvania Geological Survey, 4th ser.,
Educational Series 11, 33 p.
Konno, K., and Ohmachi, T. (1998). “Ground-motion characteristics estimated from spectral
ratio between horizontal and vertical components of microtremor.” Bull. Seis. Soc. of
America, 88 (1), 228–241.
Molnar, S., et al. (2022). “A review of the microtremor horizontal-to-vertical spectral ratio
(MHVSR) method.” J. of Seismology, pp.1-33.
Mucciarelli, M. (1998). “Reliability and applicability of nakamura’s technique using
microtremors: An experimental approach.” J. of Earthquake Eng., 2 (4), 625–638.
Nakamura, Y. (1989). “A method for dynamic characteristics estimation of subsurface using
microtremor on the ground surface.” Railway Technical Research Institute, Quarterly
Reports, 30(1), 25-33.
Nogoshi, M. (1971). “On the amplitude characteristics of microtremor, Part II.” J. Seis. Soc. of
Japan, 24, 26–40.
SESAME Project. (2004). Guidelines for the implementation of theH/V spectral ratio technique
on ambient vibrations measurements, processing and interpretation. http://sesame-
fp5.obs.ujf-grenoble.fr/Papers/HV_User_Guidelines.pdf.
Vantassel, J. (2021). jpvantassel/hvsrpy: v0. 2.1 (Version v0. 2.1) [online]. Zenodo. Available
from: [Accessed 25 May 2022].
Weary, D. J. (2015). “The cost of karst subsidence and sinkhole collapse in the United States
compared with other natural hazards.” Proc. The 14th Sinkhole Conference, Rochester,
MN.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 103

Zeng, Y., and Zhou, W. (2019). “Sinkhole remedial alternative analysis on karst lands.”
Carbonates and Evaporites, 34 (1), 159–173.
Zhou, W., and Beck, B. F. (2008). “Management and mitigation of sinkholes on karst lands: An
overview of practical applications.” Environ. Geology, 55 (4), 837–851.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 104

Towards Implementing SCPTu Geotechnical Design Guidelines for the State of Illinois

Cody Arnold1; Jorge Macedo, Ph.D., P.E.2; Paul Mayne, Ph.D., P.E.3; Luis Vergaray4;
Yumeng Zhao5; Sheng Dai, Ph.D., P.E.6; Lina Pua7; Bruce Miller8; and Brian Laningham, P.E.9
1
School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, GA.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Email: cody.arnold@gatech.edu
2
School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, GA.
Email: jorge.macedo@gatech.edu
3
School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, GA.
Email: paul.mayne@ce.gatech.edu
4
School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, GA.
Email: luis.vergaray@gatech.edu
5
School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, GA.
Email: ymzhao@gatech.edu
6
School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, GA.
Email: sheng.dai@ce.gatech.edu
7
WSP Golder Associates, Atlanta, GA. Email: lpua@golder.com
8
ConeTec, Inc., Elk Grove Village, IL. Email: bmiller@conetec.com
9
Illinois Dept. of Transportation, Georgia Institute of Technology, Springfield, IL.
Email: brian.laningham@illinois.gov

ABSTRACT

The cone penetration test (CPT) is widely used in geotechnical engineering for subsurface
soil characterization due to its nearly continuous profiling, expediency, and repeatability, which
are difficult to match with drilling, sampling, and laboratory testing, or other subsurface
characterization techniques such as the standard penetration testing (SPT). In addition to
obtaining the cone tip resistance, sleeve friction, and pore water pressure, the seismic piezocone
penetration test (SCPTu) also provides measurements of shear wave velocities with depth. A
series of SCPTu soundings have been completed at several strategic Illinois test locations to
characterize the particular response of Illinois soils in terms of stress history, strength,
compressibility, stiffness, organic content, and hydraulic properties. Additionally, since the
current geotechnical engineering practice in Illinois has a heavy reliance upon the SPT, an
evaluation of correlations between paired sets of SCPTu readings and SPT blow counts corrected
for energy efficiencies was made at a selected location. The findings of this research will be
incorporated into the Illinois Department of Transportation’s (IDOT) geotechnical manual
through the development of guidelines for using CPT in the state of Illinois. This effort is geared
towards expanding the use of CPT in IDOT practice allowing for higher quality subsurface data
that can reduce design costs while increasing sustainability and reducing risk.

INTRODUCTION

The use of penetration testing to derive subsurface soil information dates back to the late 19th
century. The most drastic change in CPT technology has been in the acquisition of data. Early
mechanical cones were only able to provide tip resistance, with sleeve friction being
incorporated some 30 years later (Sanglerat 1972). Electronic penetrometers were first developed

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 105

by the Delft Soil Mechanics Laboratory in 1948, although they were not commercially available
until the 1960s (Robertson 2001). The electric penetrometer had several advantages over the
mechanical models. These advantages included: the elimination of error due to friction between
inner and outer rods, continuous penetration at a constant rate, automated data acquisition, and
more sensitive load cells for higher accuracy in readings (Muhs 1978). The addition of the
electrical piezometer in the late 1970s allowed for the measurement of pore pressures during
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

penetration (Baligh et al. 1981, Campanella and Robertson 1981, de Ruiter 1981, Tumay et al.
1981). In 1986, the modern version of the CPT would be developed with the addition of
geophone/accelerometer. This seismic piezocone penetration test (SCPTu) was not only capable
of reading tip resistance, sleeve friction, and pore pressures but could now incorporate
measurements of shear and compression waves (Campanella et al. 1986, Robertson et al. 1986).
Given the versatility of the CPT several Departments of Transportations (DOT) in the United
States (e.g., Pennnsylvania, Minnesota, Indiana, and others) have adopted their use with details
documented in Dagger et al. (2018) and Niazi (2021) and showed the advantages of CPT testing
over other potential alternatives (e.g., the Standard Penetration Test - SPT). In this study, we
share insights from the ongoing work performed by Georgia Tech for implementing SCPTu
testing guidelines for the state of Illinois. We discuss the Georgia Tech equipment, testing
protocols, and the interpretation of representative SCPTu data collected in the state of Illinois.
The results of which will be incorporated into IDOT practice for the use of CPTu.

COMPONENTS OF THE GEORGIA TECH CPT SYSTEM

The equipment needed to perform CPT testing includes: a thrust mechanism with adequate
reaction, a data acquisition system, electric cables, push rods, and a penetrometer with desired
sensors. Figure 1a shows several of the Georgia Tech penetrometers with varying sensor
configurations, including resistivity module, u1 and u2 pore pressure sensors, seismic, and full-
size and mini-cone systems. Figure 1b shows one of several seismic sources used at Georgia
Tech for SCPTu testing. The data acquisition system consists of a signal processing unit,
transmission cable, depth recorder, and a computer. The specific components vary, but recently
developed systems contain the signal conditioning, amplification, and digital output directly
within the penetrometer as opposed to a processing unit on the surface. Modern penetrometers
will, at a minimum, provide three data channels with depth: tip resistance (qc), sleeve friction
(fs), and pore pressure (u2). The CPT group at Georgia Tech utilizes a Hogentogler SCPTu
system (a.k.a., Vertek H-T) with a 10 cm2 penetrometer that contains a dual-axis geophones for
pseudo-interval Vs measurements. The specific depth gauge used is an encoder-type depth wheel
from Hogentogler. The encoder type depth wheel has several advantages over the proximity
type. First, it is capable of higher depth resolutions (5 mm) as compared the proximity (2 cm). In
addition, it is capable of reading the push direction (i.e., up or down). This allows for more
accurate measurements as it will automatically account for any uplift in the penetrometer during
the release of the ram pressure to install new rod sections. For a more in-depth description and
discussion on the various configuration of sensors of CPT equipment and their benefits, see
Mayne (2007) and Niazi (2021).

SCPTu TESTING PROCEDURE

The CPT rig operated by Georgia Tech is shown in Figure 2. The rig consists of a Ford F-350
model with a modified flatbed. The hydraulic pushing mechanism is mounted at the rear of the

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 106

bed along with anchors and leveling pistons, as shown in Figure 2b. Due to the pushing
mechanism being on the back of the truck, it is unable to utilize the weight of the vehicle fully
and relies on the aforementioned anchors for reaction force when performing the test. The
reaction beam, as labeled, is lifted using the hydraulic ram and contains the anchor mounts. Due
to both anchors being attached to the same lifting mechanism, they must be inserted into the
ground at the same time. This can make anchoring difficult as both anchors will not always
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

penetrate at the same speed or consistency. As such, it takes a skillful operator with care to
ensure good anchor connections to the ground without inducing tilting in the mast of the ram.
Many other CPT systems will either use dead weight or individual anchors instead for efficiency
and ease of use. As important as good anchoring, leveling is also key to accurate data collection.
This CPT system also has two leveling pistons on the rear of the truck.

Figure 1. (a) Several of the Georgia Tech cone penetrometers in varying sizes and sensor
configurations, (b) One seismic source used by Georgia Tech detailed in McGillivray (2007)

Figure 2. (a) Georgia Tech CPT rig in Illinois; (b) Rear operating end of Georgia Tech’s
CPT system

The general procedure for testing with the Georgia Tech CPT truck used in the State of
Illinois consist of a site reconnaissance to determine feasible locations to perform testing. Once

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 107

the location(s) is determined, the vehicle is driven to the location and parked in a manner that the
front is slightly higher in elevation than the rear. The rear tires are blocked as well as the
emergency brake is engaged for safety but also to ensure no movement of the vehicle during the
test. From there, the rig is leveled using the pistons prior to anchoring. This is to ensure the
anchors are penetrated vertically and not at an angle. It should be noted that for SCPTu testing,
the source block is placed under one of the leveling piston feet at this point. The anchor
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

extensions and flights are then attached to the reaction beam and anchored into the ground.
Depending on the ground conditions, there are larger or smaller anchor flights. Larger flight
anchors are more difficult to anchor but provide much more reaction force. In stiff ground
conditions, for example, it may be easier and sufficient to use the smaller flight anchors.
Additionally, multiple extensions can be added to the anchors allowing for a deeper anchoring
and thus more reaction force. Typically extending the anchor depth is only necessary for very
soft soil or for deeper pushes. Once sufficient anchoring is achieved, the reaction beam is
detached from the ram and sits directly on the truck body. This allows for the ram to be raised to
advance the penetrometer while maintaining reaction force. The level of the rig is then checked
again in both length and width directions to ensure no change occurred during anchoring. Once
the rig is anchored and leveled, the electronics for the penetrometer are then assembled. Since
the rig is not enclosed, the electronics must be stored between uses and re-assembled for each
push. In terms of the shear wave velocity measurements, the seismic source used was shown in
Figure 1b. Since the particle motion is perpendicular to the propagation direction for shear
waves, the bottom plate is placed under one of the leveling piston feet parallel to the
penetrometer. This also provides a good source to ground surface contact. Additional details on
the source mechanism and source-to-receiver placement are discussed in McGillivray (2007).
Once all the components are wired and connected, the penetrometer is prepared. This
involves installing a new saturated pore pressure filter element to the penetrometer and saturating
the openings by filling the cavity with silicon grease. The tip is then installed over the filter, and
the cone is connected to the transmission wire. Once all components are ready, the software is
opened to check that the depth gauge, source, and penetrometer are all reading correctly. The
penetrometer is then ready to be attached to the friction reducer and first rod. The friction
reducer is a small attachment that is slightly larger in diameter than the rods. This is to aid in
reducing the friction from the rods as the cavity left by it is slightly larger than the rods
themselves. The penetrometer is then carefully maneuvered into the hydraulic ram, and the test
can begin. Per ASTM D 5778, the penetrometer is advanced at 2 cm/s ± 0.5 until sounding
termination. The test can be terminated for a variety of reasons: desired depth was reached,
excessive inclination, or very high tip resistance. Once the test has been completed, the software
readings are stopped, the depth wheel is removed, and the rods are pulled upward and removed
until the penetrometer is retrieved. Final baselines are then reviewed and recorded to ensure the
penetrometer load cells measurements are consistent to those at the beginning of the test.

SCPTu TESTING IN THE STATE OF ILLINOIS

Fieldwork is being conducted in the State of Illinois with the SCPTu for implementing
statewide testing guidelines. The bedrock geology of Illinois is dominated by Pennsylvanian age
rocks, which occur within the Illinois Basin. Overburden soils from alluvial and fluvial processes
lie above the bedrock. Series of SCPTu soundings have been conducted in soils at several
locations across the state to characterize the particular response of Illinois soils in terms of soil

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 108

behavior, strength, compressibility, stiffness, organic content, and hydraulic properties. Figure 3a
shows the State of Illinois with the soundings performed to date (7) and potential locations that
are being planned. The Illinois Department of Transportation (IDOT) has in total 9 districts, as
shown in Figure 3a. The SCPTu testing program considers all districts to gather a wide breadth
of data across the state. This will be particularly useful for determining the applicability of
various CPT-based screening procedures and interpretations to Illinois surficial geology. The
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

results from SCPTu, are particularly useful for examining the response of clays, which were
found in several locations, typically extending up to 10 to 15 m. Results of SCPTu tests at a
selected location (only one location is selected due to space limitations) are discussed in the
following to illustrate the type of data collected and typical responses. The site chosen was
named Florence due to its proximity to the Florence Bridge, as shown in Figure 3b.

Figure 3. (a) Location of initial SCPTu testing sites in Illinois, (b) Plan view with ConeTec
tests (from left to right; 1, 2, and 3) and (c) Site overview for CPT testing at the Florence
Bridge site

The test location is roughly 500 m from the Illinois River. The push is located within a field
on the East side of a dike that separates farmland from the river, as shown in Figure 3c.
Additional testing was performed at this area by ConeTec (See Figure 3b). Figure 4 compares
results from the SCPTu test performed by Georgia Tech and ConeTec’s soundings, considering
the first 15 m where Georgia Tech stopped the penetration as a stiff layer of gravelly sand was
hit.
Overall, the readings show good agreement between the corrected tip resistance (qt) and
sleeve friction (fs) collected data illustrating the excellent repeatability that can be achieved with
the SCPTu. Even when considering different operators and the significant distance between the
soundings. In addition to the significant distance between soundings, they have been performed
in a soybean field in which the surface conditions and soil saturation change depending on the
time of year the sounding was performed (the soundings were conducted at different times), so
seasonal related additional differences are expected, but again the general patterns are consistent.
The readings show a low qt for the upper 15 m (< 5 MPa) of material as well as low fs (< 70

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 109

kPa), which was the general pattern for the investigated Illinois clays. The water table was
determined to be located at 3 m from a dissipation test allowed to reach hydrostatic conditions.
The soil shows a positive excess pore pressure during the entirety of the push. The collected data
indicates 2 layers of clay-like soil. The upper 4 m is slightly stiffer with some silty mixtures, and
the lower portion extends to 15 m where a stiff gravelly sand layer is encountered. The CPT
charts developed by Robertson (1990, 2016), shown in Figure 5, indicate that most of the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

material behaves as “silty clay to clay” on a contractive zone with very few instances of organic
soils. The CPT charts from Schneider et al. (2008, 2012) - not shown for brevity - indicate that
the material generally behaves as a “Clay” and “Silt and Low Rigidity Index Clay” with some
instances classifying as “Sensitive Clays” and “Transitional Soils”.

Figure 4. Comparison of piezocone tests at the Florence site performed by Georgia Tech
and ConeTec Group

Figure 5. CPT Soil Behavior Type (SBTn) charts: (a) Qtn-Fr by Robertson (1990) and (b)
Update by Robertson (2016)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 110

Figure 6 shows additional geoparameters estimated from a SCPTu in the study area for
illustration purposes. The effective stress friction angle interpretation is performed using the
procedure proposed by DiBuö et al. (2019). This uses the method by Kulhawy and Mayne (1990)
for any soil classified as sand (Ic ≤ 2.6). If the soil is classified as fine-grained (Ic > 2.6), it uses
the NTH approximations from Mayne (2007). For fissured overconsolidated clays with Bq < 0.05
a modification of the NTH solution is used (Ouyang and Mayne 2019). This sounding shows the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

upper silty clay mixture (<4m) as having a friction angle between 35‫ ﹾ‬and 40‫ ﹾ‬and the underlying
clay layer having a slightly lower friction angle of 30‫ ﹾ‬to 35‫ﹾ‬. The undrained shear strength is
assessed using the Mayne and Peuchen (2018) approach. The upper clayey layer (0-4 m) shows a
decreasing strength that varies from 35 to 60 kPa. The underlying clay layer (4-15 m) shows an
increasing undrained strength between 30 and 50 kPa. The overconsolidation ratio (OCR) is
determined using the SCE-CSSM solutions by Burns and Mayne (2002). The profile appears to
be overconsolidated (OCR > 3) within the upper 4 m and has a lower OCR (< 3) for deeper soils.
The larger OCR at shallower locations may be associated with desiccation processes due to
seasonal changes and possibly farming activities (e.g., surcharge from farming equipment),
whereas the rest of the profile corresponds to a slightly overconsolidated material. The
preconsolidation stress (σp’) has been estimated considering the Agaiby & Mayne (2021)
screening procedure as 0.33 (𝑞𝑡 − 𝜎𝑣𝑜 ), 0.53(𝑢2 − 𝑢o ), and 0.6(𝑞𝑡 − 𝑢2 ), with results presented
in Figure 6. It can be noticed that 0.53(𝑢2 − 𝑢o )< 0.33 (𝑞𝑡 − 𝜎𝑣𝑜 )< 0.6(𝑞𝑡 − 𝑢2 ) for the upper 4
m as well as other instances, indicating the presence of either desiccated OC clays or organic
clays according to Agaiby & Mayne (2021). However, thermal gravimetric analyses (TGA) on
recovered samples do not suggest the presence of organics. The potential reasons for the
contrasting outcomes from the TGA and CPT-based assessments are likely due to the samples
being taken from the other side of the dike as shown in Figure 3b. Figure 6 also shows the
resulting OCR estimates using the generalized CPT procedure (Agaiby & Mayne 2021).

Figure 6. Interpretations of ’, su, OCR, and σp’ from GT SCPTu at Florence, Illinois

Figure 7a shows wave signatures at different depths, and Figure 7b shows the interpreted
shear wave velocities (Vs) from the GT sounding as well as from one of the ConeTec soundings
compared against laboratory-based measurements of Vs (i.e., using bender elements). The

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 111

consistency of the laboratory-based and CPT-based Vs estimates is remarkable for the


investigated Illinois soils, as observed in Figure 7b. The results suggest that the investigated
clays have a quite low stiffness, classifying them as soft clays based on Vs values. Figure 7c
shows the coefficient of consolidation (cvh) interpreted from dissipation tests from the sounding
west of the dike considering the CPT-based procedures from Burns and Mayne (2002) compared
against laboratory-based data (i.e., oedometer tests). The laboratory-based cvh are relatively
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

smaller compared with the CPT-based cvh estimates but still in the order of magnitude. Potential
reasons for the differences may be associated with the different boundary conditions in the field
and laboratory as well as the representability of the recovered samples.

(a) (b) (c) (d)

Figure 7. (a) Seismic wavelets throughout depth (b) Interpreted Vs of soil profile with lab
data (c) Interpreted cvh from dissipation tests and lab oedometers (d) Comparison of
measured N60 values and CPT estimated values from Georgia Tech SCPTu

Finally, given that the current geotechnical engineering practice in Illinois relies on the SPT,
the Georgia Tech CPT data were also “transformed” into equivalent SPT data using the
correlations by Lunne et al. (1997) and Robertson (2012). The Lunne et al. (1997) method was
proven to work for a wide range of soils; however, it was shown to underpredict the N60 value in
certain clays. Robertson (2012) developed an improved method considering clays. The results
are presented in Figure 7d, compared against measured SPT data. Our preliminary assessments,
shown in Figure 7d, suggest that the Robertson (2012) correlation appears to work well for the
Illinois soils, but this will be further examined as more data are collected.

CONCLUSIONS

This paper describes the initial work that has been conducted to implement statewide
guidelines on CPT for the State of Illinois. The procedures described herein are descriptive of the
Georgia Tech equipment but can be generally applied to many CPT systems. The geology of
Illinois is prime for CPT implementation since it has high applicability for assessing the
overburden clay stratum that are common soil deposits across the state. As shown in this report, a
wide range of soil parameters can be derived from a CPT sounding, including the converted N60
values. Specifically, the SCPTu sounding from the Florence site was shown to correlate well
with measured SPT N60 values using the Robertson (2012) method. Furthermore, the Vs

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 112

measurements showed consistency with laboratory-based measurements, and cvh estimates from
dissipation tests and oedometer tests are in a comparable order of magnitude. The preliminary
results related to the assessment of organic content also suggest that the Agaiby & Mayne (2021)
screening procedure for organic clays needs to be further reviewed within the context of Illinois
geology.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ACKNOWLEDGMENTS

This publication is based on the results of ICT-R27-210, Cone Penetration Testing (CPT) for
Illinois Subsurface Characterization and Geotechnical Design. ICT-R27-210 was conducted in
cooperation with the Illinois Center for Transportation; the Illinois Department of
Transportation, Office of Program Development; and the US Department of Transportation,
Federal Highway Administration.
The authors would also like to thank ConeTec, Inc. of Chicago, Illinois for their generous in-
kind support to this project.

REFERENCES

Agaiby, S. S., and Mayne, P. W. (2021). CPTU identification of regular, sensitive, and organic
clays towards evaluating yield stress profiles. AIMS GeoSciences 7 (4): 553-573.
Baligh, M. M., Azzouz, A. S., Wissa, A. Z. E., Martin, R. T., and Morrison, M. J. (1981). The
piezocone penetrometer. Cone Penetration Testing and Experience. American Society of
Civil Engineers, Reston, VA: 247-263.
Burns, S. B., and Mayne, P. W. (2002). Analytical cavity expansion-critical state model for
piezocone dissipation in fine-grained soils. Soils and Foundations 42 (2): 131–137.
Campanella, R. G., and Robertson, P. K. (1981). Applied cone research. Cone Penetration
Testing and Experience. ASCE, Reston, VA: 343–362.
Campanella, R. G., Robertson, P. K., and Gillespie, D. (1986). Seismic cone penetration test. Use
of In-Situ Tests in Geotechnical Engineering, Geotechnical Special Publication 6, ASCE,
Reston, VA: 116–130.
Dagger, R., Saftner, D., and Mayne, P. W. (2018). Cone Penetration Test Design Guide for State
Geotechnical Engineers. Minnesota Department of Transportation. Retrieved from the
University of Minnesota Digital Conservancy, https://hdl.handle.net/11299/203697.
DiBuö, B., D’Ignazio, M., Selãnpaã, J., Länsivaara, T., and Mayne, P. W. (2019). Yield stress
evaluation of Finnish clays based on analytical CPTu models. Canadian Geotechnical
Journal 57 (11): 1623-1638.
de Ruiter, J. (1981). Current penetrometer practice: Cone Penetration Testing and Experience.
ASCE, Reston, VA: 1–48.
Kulhawy, F. H., and Mayne, P. W. (1990). Manual on Estimating Soil Properties for Foundation
Design. Report EL-6800, Electric Power Research Institute, Palo Alto: 306 p.
Liao, T., and Mayne, P. W. (2006). Automated post-processing of shear wave signals. Proc. 8th
US National Conference on Earthquake Engineering, San Francisco, pp. 460.1-460.10.
Lunne, T., Robertson, P. K., and Powell, J. J. M. (1997). Cone Penetration Testing in
Geotechnical Practice. EF Spon/Blackie Academic, London: 312 p.
Mayne, P. W. (2005). Integrated ground behavior: In-situ and lab tests. Deformation
Characteristics of Geomaterials, Vol. 2: Taylor & Francis Group, London: 155–177.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 113

Mayne, P. W. (2007). NCHRP Synthesis 368 on Cone Penetration Test. Transportation Research
Board, National Coop. Highway Res. Program, Washington, DC: 120 p.
Mayne, P. W., and Peuchen, J. (2018). CPTu bearing factor Nkt for undrained strength
evaluation in clays. Proceedings of The Fourth International Symposium on Cone
Penetration Testing (CPT 2018, Delft) CRC Press, London.
McGillivray, A. V. (2007). Enhanced integration of shear wave velocity profiling in direct -push
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

site characterization systems. PhD dissertation, Civil & Environmental Engineering, Georgia
Inst. of Technology, Atlanta, GA.
Muhs, H. (1978). 50 years of deep soundings with static penetrometers. Berlin Universita¨ t,
Deutsche Forschungsgesellschaft fur Boden-Mechanik (Degebo). Mitteilungen, 34, 45–50.
Niazi, F. (2021). CPT-Based Geotechnical Design Manual, Volume 1: CPT Interpretation—
Estimation of Soil Properties. FHWA/IN/JTRP-2021/22. Indiana Dept. of Transportation.
Ouyang, Z., and Mayne, P. W. (2019). “Modified NTH method for assessing effective friction
angle of normally consolidated and overconsolidated clays from piezocone tests.” Journal of
Geotechnical and Geoenvironmental Engineering 145(10):04019067.
Robertson, P. K., Campanella, R. G., Gillespie, D., and Grieg, J. (1986). Use of piezometer cone
data. Use of In-Situ Testings in Geotechnical Engineering (Geotechnical Special Publication
No. 6), ASCE, Reston, VA: 1263–1280.
Robertson, P. K. (1990). Soil classification using the cone penetration test. Canadian
Geotechnical Journal 27(1), 151–158.
Robertson, P. K., and Wride, C. (1998). Evaluating cyclic liquefaction potential using the cone
penetration test. Canadian Geotechnical Journal 35, 442-459.
Robertson, P. K. (2001). Sixty years of the CPT—How far have we come? Proceedings,
International Conference on In-Situ Measurement of Soil Properties and Case Histories,
Bali, Indonesia: 1–16. PDF at www.usucger.org.
Robertson, P. K. (2012). Mitchell lecture. Interpretation of in-situ tests—Some insight.
Geotechnical & Geophysical Site Characterization, Vol. 1 (Proc. ISC-4, Pernambuco),
Millpress, Rotterdam: 3–24.
Robertson, P. K. (2016). CPT-based soil behaviour type (SBT) classification system—an update.
Canadian Geotechnical Journal, 53(12), 1910–1927.
Sanglerat, G. (1972). The Penetrometer and Soil Exploration. Elsevier Publishing Company,
Amsterdam: 464 p.
Schneider, J. A., Randolph, M. F., Mayne, P. W., and Ramsey, N. R. (2008). Analysis of factors
influencing soil classification using normalized piezocone tip resistance and pore pressure
parameters. Journal of Geotechnical and Geoenvironmental Engineering, 134(11), 1569–
1586.
Schneider, J. A., Hotstream, J. N., Mayne, P. W., and Randolph, M. F. (2012). Comparing CPTU
Q-F and u2/v0' soil classification charts. Geotechnique Letters, 2(4), 209–215.
Tumay, M. T., Boggess, R. L., and Acar, Y. (1981). Subsurface investigation with piezo-cone-
penetrometer. Cone Penetration Testing and Experience. ASCE National Convention, St.
Louis. American Society of Engineers, Reston, VA: 325-342.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 114

Determination of Geotechnical Properties in Intermediate Geo-Materials with Newly


Developed In Situ Test Device

Young-Woo Song, S.M.ASCE1; Jung-Hyun Ryu2; Inhyun Kim3;


and Choong-Ki Chung, Ph.D., M.ASCE4
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Ph.D. Candidate, Dept. of Civil and Environmental Engineering, Seoul National Univ., Seoul,
Republic of Korea. Email: songyw@snu.ac.kr
2
Ph.D. Student, Dept. of Civil and Environmental Engineering, Seoul National Univ., Seoul,
Republic of Korea. Email: ryoujh1155@snu.ac.kr
3
Ph.D. Student, Dept. of Civil and Environmental Engineering, Seoul National Univ., Seoul,
Republic of Korea. Email: hyunin1384@snu.ac.kr
4
Professor, Dept. of Civil and Environmental Engineering, Seoul National Univ., Seoul,
Republic of Korea (corresponding author). Email: geolabs@snu.ac.kr

ABSTRACT

Intermediate geomaterials (IGMs) are used as bearing strata in various civil engineering
projects. It is important to reliably evaluate the geotechnical characteristics of IGMs to ensure
the stability of their associated structures. In the case of IGMs, such as highly or completely
weathered rocks and residual soils, it is difficult to obtain undisturbed specimens for laboratory
tests. Therefore, field tests are performed to evaluate the design parameters of IGMs. However, it
should be noted that conventional field tests are specialized for soils and rocks. In this study, a
novel field test device, the borehole pressure shear tester (BPST), was developed to reliably
evaluate the deformation and shear strength parameters of IGMs. A 1:1 scale physical model test
was conducted to validate the reliability and applicability of the BPST. The test was performed
using a calibration chamber, for simulating the stress conditions of the field ground. The test
results were compared with laboratory triaxial compression test results and direct shear test
results.

INTRODUCTION

Intermediate geomaterials (IGMs) are transitional materials between soil and rock in terms of
their strength and compressibility; examples include residual soils and very weak rocks
(AASHTO, 2020). IGMs are distributed across various regions. Several states in the United
States, such as Wyoming, Kansas, Idaho, and Montana, feature extensive IGM distributions
(Islam, 2021). In regions with geological features where IGMs are widely distributed, they are
mainly used as bearing strata for structures in various civil engineering projects that require pile
foundations. Therefore, it is important to reliably evaluate the geotechnical properties or design
parameters of IGMs, for ensuring the stability of the associated structures.
Design parameters can be most accurately evaluated using a load test; however, owing to the
cost and time constraints, design parameters are typically evaluated in laboratory tests or by
performing simple in situ tests. However, in the case of IGMs such as residual soils or weak
weathered rocks, such as highly or completely weathered rocks, it is very difficult to obtain
undisturbed specimens suitable for laboratory tests. Therefore, it is desirable to evaluate the
design parameters through in-situ tests. Although various design parameters are required for

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 115

designing ground structures, it is important to reasonably determine the deformation and shear
strength parameters from the perspectives of bearing capacity and settlement evaluation. Field
tests, including the standard penetration test (SPT), pressuremeter test (PMT), and borehole shear
test (BST), are performed for evaluating the deformation and shear strength parameters of IGMs.
However, general field tests are specialized for soils and rocks. Therefore, it is inappropriate to
apply conventional field tests to IGMs.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

In detail, it has been known that design parameters can be estimated based on the SPT.
Various researchers have conducted studies on the correlation between SPT-N values and
deformation or shear strength parameters (Anbazhagan et al., 2012; Briaud, 2013; Peck et al.
1974). However, the reliability of the design parameters estimated based on the SPT is low,
owing to the excessive blow count on IGMs such as dense residual soils and soft rocks (Lee et al.
2019; Benoit and Howie, 2014). On the other hand, the PMT is often used for estimating the
deformation parameters of unusual geomaterials such as IGMs, and its reliability is relatively
high. However, complicated calculations with several assumptions are required for evaluating
the shear strength parameters (Haberfield and Johnston, 1990). Meanwhile, the BST and the rock
BST (RBST) have been used to directly evaluate the shear strength parameters of geomaterials in
the field. However, it should be noted that both the BST and RBST have the potential to
misestimate the shear strength of IGMs, because they have been developed for soft-medium soils
and rocks, respectively (Mitchell, 1979; Handy and Fox, 1967; Handy et al., 1976). The shape of
a shear plate and the loading capacity of the BST and RBST are not for IGMs. Furthermore, they
were designed not to measure the normal displacement and have limitations in measuring normal
displacement in an analog manner.
In this study, a novel field test device, the borehole pressure shear tester (BPST), was
developed to reliably evaluate the deformation and shear strength parameters of IGMs using the
same device. The BPST was designed to generate the horizontal load-displacement curve and
vertical load-displacement curve. To achieve this purpose, the BPST consists of a lower part
inserted into the borehole and an upper part mounted on the borehole. The lower part is in charge
of the horizontal loading system and the upper part is in the charge of the vertical loading
system. They include a load cell and encoder motor respectively to measure the horizontal and
vertical stresses and accompanying displacements. Therefore, the deformation parameters can be
evaluated by applying a horizontal force to the ground, and the shear strength parameters can be
evaluated by shearing the ground. A 1:1 scale physical model test was conducted to validate the
reliability and applicability of the BPST. The tests were performed using a calibration chamber
to simulate the stress conditions of the field ground. The deformation and shear strength
parameters were evaluated based on the tests. Finally, the results were compared with laboratory
triaxial compression test and direct shear test results.

BOREHOLE PRESSURE SHEAR TESTER

Configuration

To evaluate the deformation and shear strength parameters, the reaction force should be
measured by applying pressure to the ground and the shear strength should be measured by
shearing the ground. The BPST was developed to perform the aforementioned functions, such as
the pressure and shear tests. The configuration of the BPST is shown in Figure 1. It consists of an
upper section and a lower section. Each section has a pressurization system with a direct current
(DC) encoder motor, to ensure that the pressurization capacity is adequate for IGMs.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 116

The function of the lower section is related to the pressure test. This section includes shear
plates, an encoder motor, a load cell, and a wedge-shaped load transferer. The encoder motor
makes it possible to push the shear plate to the borehole wall and to measure the horizontal
displacement based on the encoding value. The load cell is used for measuring horizontal loads.
The function of the upper section is related to the shear test. This section includes a control
box, a load cell, and a rod. The upper box includes an encoder motor, gears, and a reaction plate.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

The encoder motor makes it possible to pull the lower section upward and measure the vertical
displacement. The load cell is used for measuring vertical loads.

(a) (b)

Figure 1. Configuration of the BPST: (a) upper part and (b) lower part

The shape of the shear plate is an important aspect of the shear test. Various studies have
been conducted, and the results of these studies have shown that the number and spacing of
wedges should be adjusted according to the target geomaterial. Conventional BST has twenty-
five continuous wedges on each shear plate and its length is 6.35cm (Figure 2(a)). In contrast, an
RBST has two separated wedges on each shear plate (length, 2.2 cm). This is because narrow
spacing and a large number of wedges result in an inadequate seating of the shear plate on hard
soils and rocks, causing the shear strength misestimation. Meanwhile, it should be noted that
IGMs have different properties from those of soils or rocks. Therefore, the BPST was developed
by considering the optimal shear plate shape for IGMs. A laboratory test was conducted to
simulate the in situ BST, for determining the optimal shape of the shear plate. The determined
shape is shown in Figure 2(b). There are three wedges, with the inter-wedge spacing of 15 mm,
amounting to five times the wedge length.

Loading and measuring system

The BPST has two loading systems: horizontal and vertical. The horizontal-loading system of
the BPST is shown in Figure 3(a). When power is supplied by the encoder motor, the rod
connected to the lower part is pushed in the vertical direction, the wedge-shaped load transfer
connected to the rod moves downward, and the shear plates bound to both sides move
horizontally along the rail. Therefore, a vertical force is transmitted in the horizontal direction by
a wedge-shaped member. Consequently, a horizontal force is applied to the shear plates. The
horizontal load applied to the shear plates can be directly measured by the load cells inside each
shear plate, and the displacement can be measured using the encoding value of the encoder
motor.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 117
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Shear plates for (a) the conventional BST and (b) the BPST ((a) was retrieved
from Handy Geotechnical Instruments (n.d.))

The vertical-loading system is shown in Figure 3(b). In the vertical-loading system, a load
can be applied through the encoder motor, and the displacement can be measured from the
encoding value in the same manner as in the horizontal-loading system. When power is supplied
from the encoder motor, the rod connected to the lower part moves in the vertical direction, and
the shear head connected to the rod is pulled upward and shears the ground in the borehole. The
applied vertical load can be measured using the load cell between the upper loading part and
connecting rod.

Figure 3. Loading and measuring systems of the BPST: (a) horizontal and (b) vertical

EXPERIMENTAL METHODS

Experimental equipment

The BPST is a field test equipment developed to be used for the ground investigation and its
lower section was designed to be placed inside the borehole, similar to the other in situ tests

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 118

(PMT, BST). The in situ tests reflect the characteristics of the field site, such as the confining
load according to the groundwater level and depth. Therefore, to validate the BPST, it was
necessary to conduct tests under conditions similar to the actual field conditions. In particular,
because the confining stress of the ground strongly affects the ground characteristics, a
calibration soil chamber, as shown in Figure 4, was used for simulating the field site stress
conditions. The calibration soil chamber applied vertical pressure to the ground by applying
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

pneumatic pressure to the loading plate using an air compressor. In addition, for simulating the
shape of the field borehole in model tests, an NX-size hole casing was placed inside the
calibration soil chamber, and the ground was prepared only around the casing.

Figure 4. Calibration soil chamber

Experimental material and condition

Residual soil is a geomaterial that is commonly distributed in areas with severe weathering,
owing to its geological characteristics, and is classified as an IGM. Therefore, residual soil was
used for the tests in this study. The tested residual soil was obtained from a field in South Korea,
and the index properties were determined based on laboratory tests, as shown in Table 1.

Table 1. Index properties of the residual soil sample

USCS classification SW Specific gravity, Gs 2.63


Minimum dry unit weight, γ𝑑,𝑚𝑖𝑛 Maximum dry unit weight, γ𝑑,𝑚𝑎𝑥
12.6 18.8
(kN/m3) (kN/m3)
Coefficient of uniformity, C𝑢 7.9 Coefficient of curvature, Cc 1.55
Optimal water content, w𝑜𝑝𝑡 (%) 12.2 Percent finer than #200 sieve (%) 3.59

IGMs can be defined as materials with corrected SPT N-values in the range of 50–100 blows
per 0.3 m (Martin and Stacey, 2018; O'Neill and Reese, 1999). Terzhagi and Peck (1948)
presented the correlation between N-values, relative density, and friction angle. According to
them, the relative density is more than 85% for N-values above 50. Therefore, residual soil was

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 119

prepared with a relative density of 90% for the tests using rammer compaction. The vertical
confining pressures of 80kPa and 160kPa were adopted for tests. They correspond to the vertical
stress of the unsaturated field ground at about 4m and 8m depth, respectively.

Experimental procedure
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Before the application of the new device, the laboratory shear tests (direct shear test, triaxial
compression test) were conducted to compare the results with that of the physical model test. The
tests were conducted using the same test material under the same test conditions. The test results
are shown in Figure 5.
As mentioned in the configuration chapter, the BPST performs two important tests of
pressure and shear tests without changing the device. Therefore, the tests were conducted
successively after preparing the ground and pulling out the NX-size casing.
First, the pressure test was conducted by placing the shear plate of the shear head at the test
position and applying a horizontal load to the borehole wall. The pressure test was performed as
a strain-control test. During the tests, the load and displacement were measured every 1 second.
Consequently, the horizontal load-displacement behavior was evaluated.
Second, the shear test was performed by expanding the shear plate in the horizontal direction,
penetrating it into the ground, and then pulling out the probe upward. In this process, the ground
between the wedges of the shear plate was sheared and the shear stress and displacement were
measured.

(a) (b)

Figure 5. Test results for (a) direct shear test and (b) triaxial compression test

EXPERIMENTAL RESULTS

Pressure test

The pressure on the ground and the ground displacement were measured based on the pressure
test. The relationship between stress and strain was derived, as shown in Figure 6. In the stress-

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 120

strain curve, three zones were identified: 1) a non-contact and disturbed zone, 2) a pseudo-elastic
zone, and 3) a plastic zone. This partitioning was similar to the results of the PMT. The strain
value, which is the standard for the classification of two zones (the non-contact and disturbed
zone and the pseudo-elastic zone), was 6.9%, which is the strain required until the wedge height
of the shear plate penetrates the ground completely. The modulus of deformation, which is a
design parameter, could be evaluated using the slope in the pseudo-elastic zone.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 6. Horizontal stress-strain curve

The pressuremeter modulus is a parameter taken from PMT. Although the BPST has a finite
curve-shaped pressurizing unit, it covers a wide area when it pressures the borehole wall. Thus,
the modulus of deformation was determined using Eq. (1) similarly to PMT. The modulus of
deformation was determined using the data for strains in the 6.9–12% range, where the linearity
is well maintained. To evaluate the appropriateness of the determined modulus, it was compared
with the secant modulus E50 of the triaxial compression test. The comparison results are
presented in Table 2. The modulus of deformation evaluated by the BPST was similar to the
secant modulus E50 of the triaxial compression test. The evaluated modulus was normalized to
the vertical confining pressure, and the results showed similar values. Thus, it was concluded
that the BPST can be used for estimating the deformation moduli of residual soils, regardless of
the confining pressure.
∆𝑃
𝐸𝑚 = (1 + 𝜐) ∙ 𝑟 ∙ ∆𝑅 (1)

In the above, Em is the deformation modulus, υ is the Poisson ratio, r is the intermediate
radius for the section used to calculate the slope, ΔP is the change in stress, and ΔR is the change
in radius.

Shear test

As a result of the shear test, the relationship between the shear stress and displacement was
derived, and the results are shown in Figures 7(a) and (b). Figure 7(a) shows the results under

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 121

80kPa vertical stress and various normal stresses of 78kPa, 148kPa, 163kPa, and 189kPa.
Figures 7(b) shows the results under 160kPa vertical stress and various normal stresses of
101kPa, 166kPa, 193kPa, and 233kPa. It should be noted that the shear stress increased as the
normal stress increased. The shear stress-displacement behavior exhibited a trend similar to that
of the hardening behavior. Therefore, in this study, a displacement of 3 mm was assumed as the
failure point because the shear stress seemed to converge at this point. The shear strength was
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

evaluated based on this assumption, and a failure envelope was drawn, as shown in Figure 8,
assuming that the relative density change is small. The shear strength parameters were derived
from the failure envelope and compared with the shear strength parameters determined from the
DST. The results are presented in Table 3.

Table 2. Deformation modulus evaluated from BPST and TXT

Relative Vertical
Triaxial compression Physical model Physical model
density confining
test, 𝐸50 (MPa) test, 𝐸𝑚 (MPa) test, 𝐸𝑚 /√𝜎
(%) stress (kPa)
80 10.7 9.6 1.07
90
160 18.0 14.2 1.12

(a) (b)

Figure 7. Shear stress-displacement curve for vertical stress of (a) 80kPa and (b) 160kPa

Since the test material is well-graded sand, it may be desirable to analyze the test results
assuming the cohesion to be zero. It may be also desirable to assume that the cohesion is non-
zero because the DST results showed that the cohesion value is too large to be neglected.
However, it should be noted that the cohesion estimated from the test is affected by the normal
stress condition due to the non-linear characteristics of the material. Thus, it is desirable to

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 122

directly compare the shear strengths measured in DST and the physical model test. In this case,
they showed similar shear strength values under similar normal stress conditions. Also, it can be
observed that the shear strength parameters determined from the BPST were similar to those
determined from the DST when cohesion was assumed to be zero.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 8. Failure envelope

Table 3. Shear strength parameters evaluated from BPST and DST

Direct shear test Physical model test (BPST)


Relative
Cohesion Friction angle Cohesion Friction angle
density (%)
(kPa) (°) (kPa) (°)
C=0 0 52.2 0 51.2
90
C≠0 42.6 43.0 3.3 50.7

CONCLUSION

IGMs differ from soils and rocks in their characteristics. These materials can be used as
bearing strata for geotechnical structures; therefore, it is necessary to determine the design
parameters of IGMs. However, an appropriate test method for evaluating the design parameters
of IGMs has not yet been developed. In this study, a 1:1 scale physical model test was performed
using a novel field-test device, which was developed and specialized for IGMs. Physical model
tests included pressure and shear tests. The test results were compared with the results of the
triaxial compression and direct shear tests. The horizontal stress-strain curve was obtained from
the pressure test, revealing three zones: 1) the non-contact and disturbed zone, 2) the pseudo-
elastic zone, and 3) the plastic zone. This behavior was similar to the behavior that is obtained
using the PMT. The soil modulus was estimated from the pseudo-elastic zone, yielding a value
similar to the E50 value for the TXT. The shear stress-strain curve was obtained from the shear

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 123

test, and the curve showed that the shear stress increased with increasing normal stress. The
shear strength parameters were determined from the failure envelope drawn using the peak shear
stress, and they were similar to those of the DST.

ACKNOWLEDGMENT
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

This research was supported by a grant (21SCIP-B119958-06) from the Smart Civil
Infrastructure Research Program funded by the Ministry of Land, Infrastructure, and Transport of
the South Korean government and by the Institute of Construction and Environmental
Engineering of Seoul National University. The authors wish to express their gratitude for this
support.

REFERENCES

AASHTO. (2020). AASHTO LRFD Bridge Design Specifications 9th ed. U.S. Customary Units.
American Association of State Highway and Transportation Officials, Washington, D.C.
Anbazhagan, P., Parihar, A., and Rashmi, H. N. (2012). Review of correlations between SPT N
and shear modulus: A new correlation applicable to any region. Soil Dynamics and
Earthquake Engineering. 36. 52-69.
Benoît, J., and Howie, J. A. (2014). A view of pressuremeter testing in North America. Soils
Rocks. 37, 211–231.
Briaud, J. L. (2013). Geotechnical engineering: unsaturated and saturated soils. John Wiley &
Sons.
Haberfield, C. M., and Johnston, I. W. (1990). The interpretation of pressuremeter tests in weak
rock-theoretical analysis. In: Pressuremeter. Proceedings of the Third International
Symposium on Pressuremeters. 169–178.
Handy Geotehnical Instruments. (n.d.). Borehole shear tester. Retrieved October 13, 2022, from
https://handygeotech.com/borehole-shear.
Handy, R. L., and Fox, N. X. (1967). A soil borehole direct shear test device. Highway Research
News. 27, 42–52.
Handy, R. L., Pitt, J. M., Engle, L. E., and Klockow, D. E. (1976). Rock borehole shear test.
Proceedings of the 17th U.S. Symposium on Rock Mechanics. 1–11.
Islam, M. S. (2021). Pile behaviors in shales through full-scale static load testing, dynamic
testing, and finite element analysis. Master thesis, Department of Civil and Architectural
Engineering, University of Wyoming, Laramie, WY.
Lee, S. H., Baek, S. H., Song, Y. W., and Chung, C. K. (2019). Case study of correlation
between the SPT- N Value and PMT results performed on weathered granite zone in Korea.
J. Korean Geotech. Soc. 35, 15–24.
Martin, D., and Stacey, P. (2018). Guidelines for open pit slope design in weak rocks. Australia:
CSIRO
Mitchell, J. K. (1979). “In Situ Techniques for Site Characterization,” Site Characterization and
Exploration, American Society of Civil Engineers, New York, pp. 107-129
O’Neill, M. W., and Reese, L. C. (1999). Drilled shafts: Construction procedures and design
methods. Publication No. FHWA-IF-99-025.McLean, VA: Federal Highway Administration,
Office of Implementation.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 124

Peck, R. B., Hanson, W. E., and Thornburn, T. H. (1974). Foundation engineering, 2nd Ed.,
Wiley, New York.
Terzaghi, K., and Peck, R. B. (1948). Soil Mechanics in Engineering Practice. John Wiley.
Yang, H., White, D. J., and Schaefer, V. R. (2006), “In-situ Borehole Shear Test and Rock
Borehole Shear Test for Slope Investigation”, In Sit\ Geomaterial Characterization, pp.293-
298.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 125

Application of Geostatistical Sequential Simulation Methods for Probabilistic 3D Subsoil


Modeling and Uncertainty Quantification Concept and Examples

Andreas Witty1; Andrés A. Peña-Olarte2; and Roberto Cudmani3


1
Chair of Soil Mechanics and Foundation Engineering, Rock Mechanics and Tunneling,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Technical Univ. of Munich, Munich. Email: a.witty@tum.de


2
Chair of Soil Mechanics and Foundation Engineering, Rock Mechanics and Tunneling,
Technical Univ. of Munich, Munich. Email: a.pena@tum.de
3
Chair of Soil Mechanics and Foundation Engineering, Rock Mechanics and Tunneling,
Technical Univ. of Munich, Munich. Email: r.cudmani@tum.de

ABSTRACT

The application of geostatistical sequential simulation methods to build geomechanical


models and quantify the corresponding uncertainty is well established in the oil and gas industry,
whereas the majority of subsoil models created for civil engineering projects utilize deterministic
interpolation methods where uncertainty cannot be evaluated. In this paper, the root causes of
these uncertainties for subsoil model generation are briefly explained, and a concept for the
consideration of these uncertainties by means of probabilistic subsoil modeling is presented. This
makes it possible to statistically evaluate geotechnical risks in both the design and construction.
Using two case studies, the use of the geostatistical method is demonstrated on the basis of site
investigations consisting of (1) CPTs and (2) borehole logs. The case studies illustrate the great
potential and advantages of this methodology of subsoil modeling compared to the conventional
deterministic approach. Further developments are required for its implementation in practice,
which are outlined in a brief outlook.

INTRODUCTION

The task of engineering geologists and geotechnical engineers is to characterize and model
the spatial arrangement of the soil layers and to quantify and evaluate the uncertainties of such
arrangement with regard to the planning and execution of construction projects. Since a spatial
3D subsoil model is developed by, among other things, interpolating the results of punctual
investigations (exploratory drillings, field and laboratory tests), it is always subject to
uncertainties. A distinction must be made between uncertainties in the subsoil structure and
uncertainties in the soil properties. Uncertainty in subsoil structure relates to the geometry of
geological features and can be modeled by geostatistical methods. Uncertainty in soil properties
relates to the expected distribution of soil properties within soil layers and can be quantified
using data-based methods.
The purpose of this paper is to show a path to the digital future of geotechnical 3D subsoil
modeling by considering the associated uncertainties. Current advances in digitization and the
establishment of data exchange formats make it easy to create data banks, as well as develop
data-based or implicit modeling approaches and analyses in the field of engineering. In implicit
soil modeling, layer boundaries, for example, are not drawn by hand (explicitly), but are
automatically interpreted from geodetic and soil exploration data. This enables the addition of
new data at any time and facilitates the straightforward adjustment and refinement of the models.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 126

For the application of these methods, so called voxel (volume pixel) or block models, in
which the subsoil is subdivided into pixels or into 3D voxel blocks similar to a 3D digital image,
are clearly more advantageous than 2 or 2.5 D layer models, in which the subsoil is subdivided in
layers. Information on subsoil composition and properties, as well as the corresponding
uncertainty, can be assigned to these blocks. The methodology described herein enables a
probabilistic viewing of the subsoil by calculating probabilities for the occurrence of a certain
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

geological feature and its properties.


In this paper, the terminology and the concept of probabilistic subsoil modeling are first
explained. Then the practical implementation of the described concept for modeling spatial
uncertainties by sequential simulation is exemplified in two case studies. In particular, the digital
spatial representation of soil layers as a block or voxel model is discussed.

UNCERTAINTIES IN THE SPATIAL DISTRIBUTION OF SOIL LAYERS AND


ASSOCIATED PROPERTIES.

For the purpose of planning and design, the subsoil is divided into soil layers or
homogeneous zones with similar geological and geotechnical properties. For the representation
of uncertainty in the model, it is proposed to make a distinction between parametric and spatial
uncertainty (Wang et al. 2020). The objective of probabilistic subsoil modeling is to analyze and
represent the spatial variability of the layered structure and the parametric variability of the
associated physical and geotechnical properties (see Figure 1). This approach to quantify the
uncertainties allows clearer communication of the state of knowledge and better management of
risks in the design and construction process.

Figure 1: Process of creation of a probabilistic 3D subsoil model considering uncertainty


related to the spatial extent of soil layers discretized in voxel blocks and soil properties
(modified after Wang et al. 2020).

In contrast to modeling soil layers as surfaces, 3D block models allow spatial modeling of
subsoil properties and the related uncertainties. Parametric values from field and laboratory tests
are assigned to the soil layers. According to Ching and Phoon (2020), subsoil investigation data
can be characterized by the following attributes: multivariate, uncertain, insufficient, site-
specific, incomplete and erroneous. Methods such as Bayesian machine learning can be used to
derive confidence intervals and characteristic values from such data. Interpolation methods are
not sufficient for quantification of spatial uncertainty, but rather only simulation methods are
appropriate (e.g. Pyrcz and Deutsch 2014, Grasmick 2019). Probabilistic modeling aims to create
a set of models that represent the probability space.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 127

The Sequential Gaussian Simulation (SGSIM) of random fields or realizations enables the
capture of the heterogeneity of the ground, based on the known data points and the simulation of
the relevant ground properties. Values are calculated for the simulation for all blocks in the
model in random order in succession using Kriging and a random sample is drawn from the
distribution with the local Krige variance. If many realizations are calculated, the mean value
approaches the estimate of simple kriging. A frequency or probability distribution of the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

corresponding soil property can be calculated for each block (Pyrcz and Deutsch 2014).
Spatially located variables or data of the same category that are spatially distributed in the
model area and available in large numbers are suitable for determining the correlation lengths. A
good prerequisite is provided by Standard Penetration Tests (SPT) or Cone Penetration Tests
(CPT), but also by exploratory drillings.
Sequential Gaussian Simulation (SGSIM) is suitable for continuous variables, e.g., tip
resistance of CPTs. Griffith et al. (2009), Lloret-Cabot et al. (2014), or Yang et al. (2017) applied
conditional simulation for the spatial modeling of CPT data and analysing slope stabilities. This
approach is presented in the first case study and applied to soil classification based on Robertson
(2010).
Besides modeling the layers separately, sequential simulation methods can be used for
categorical variables, e.g., layers in a borehole description of a geologist. Grasmick (2019) used
Sequential Indicator Simulation (SISIM) and Plurrigaussian Simulation (PGSIM) for a tunnel
project to predict settlements. Gangrade et al. (2022) used PGSIM for the prediction of mixed
face condition for a tunnel constructed by a tunnel boring machine. Kring and Chatterjee (2020)
have combined the use of SISIM and SGSIM for slope stability analysis modeling the Rock
Quality Designation values as continuous variable and fault zones as categories. An approach for
the combination of SGSIM and SISIM for the modeling of a sedimentation and erosion cycle is
presented in the second case study.

CASE STUDY 1: CONE PENETRATION TESTS IN ALLUVIAL SEDIMENTS

In the first case study, a 2D model of the spatial distribution the Soil Behavior Type (SPT)
after Robertson (2010) was derived from cone resistance and sleeve friction of seven CPTs
located in a pre-alpine valley filled with alluvial gravel, sand and clay sediments. It shows how
the spatially distributed uncertainty of a given soil type can be modeled and the advantages of
statistical simulation (SGSIM) compared to interpolation methods. The realizations were
simulated with the open-source tool, GeostatsPy (2021). The process is depicted in Figure 2 and
was applied to sleeve friction fs and the corrected cone resistance qt. During post-processing, the
realizations were used to calculate the SBT and quantify the uncertainty of the classification.
This independent modeling is valid because qt and fs did not show a correlation, otherwise, Co-
Simulation would be appropriate.
Firstly, the data must be trend-adjusted and transformed into a standard normal distribution.
Cone Penetration Tests, for example, often have a vertical trend due to the influence of pressure
on cone resistance which must be removed. The CPT values did not show a trend that needed to
be corrected, and the logarithm of the sleeve friction fs and the corrected cone resistance qt was
directly transformed into the standard normal distribution. To reduce the computational effort,
the data was reduced to a resolution of 0.1 m in depth.
Directional geological processes (e.g., sedimentation, erosion) result in anisotropic bedding
conditions, so therefore two experimental semivariograms are needed in the horizontal direction

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 128

(longitudinal/transverse) and one in the vertical direction. The semivariogram represents the
variance of two values as a function of spatial distance. The experimental semivariogram is
calculated from the data by computing the semivariances from all data points for various
distances. In order to simulate a value at all arbitrary points in the model space, a theoretical
semivariogram model defined by a mathematical function (e.g., spherical, exponential or
Gaussian function) must be fitted to the experimental semivariogram. Now a grid is defined, the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

points of which correspond to the centers of the individual blocks in the block model. The size of
the individual blocks is chosen based on the semivariogram model as well as the practical
relevance. Blocks that are too large lead to an overgeneralization of the soil conditions, and
blocks that are too small require large computational capacities and memory. Once a block
model has been created with the geometric parameters, semivariogram-based simulation is
performed to quantify the uncertainty. The algorithm therefore visits every block, interpolates the
value using previously known and simulated data using Kriging interpolation and then adds a
random value drawn from a distribution defined by the Kriging variance. For details of the
simulation algorithm see Deutsch and Journel (1998).

Figure 2: Flow chart for the generation of SGSIM realizations.

With increasing number of realizations, the spatial distribution of the mean values becomes
smoother and the estimation of the uncertainty more accurate. For the case study 500 realizations
were calculated. The results where then backtransformed and in this case no trend needed to be
added. During postprocessing, arbitrary realizations of qt and fs were combined to calculate
realizations of the Soil Behavior Type Index ISBT, which was then classified according to
Robertson (2010) to determine the most probable SBT from the realizations and to calculate the
entropy.
In addition to the variance, the information entropy (Shannon 1948) is a suitable and
frequently used measure to represent the reliability of the prediction for a given homogeneous
zone. The entropy H is calculated as the sum of the products of the probability of occurrence pi
of the N soil layers and its logarithm:

H = – ΣiN(pi × log(pi)) (1)

A low entropy corresponds to a low uncertainty regarding the classification of the soil in a
particular soil layer. High entropy values indicate that the classification is ambiguous and the
prediction is uncertain. The calculated entropy for the example is shown in Figure 3b.
In Figure 3a, the friction ratio, i.e., the ratio of the sleeve friction and the tip resistance, was
used to classify the soil into soil types with similar geotechnical behavior according to Robertson

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 129

(2010). The probability that a block must be assigned to a different soil behavior type can now be
calculated using the frequency distribution of every cell.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3: a) Most probable soil behavior type after 500 simulations; b) Corrected tip
resistance qt and uncertainty displayed as entropy related to the most probable soil type.

The model in Figure 3 shows the central tendency of the Gaussian distribution in this case.
As can be seen from the predicted soil type, the fraction of extreme data values is underestimated
by the approach. A transformation in categorical indicators or prior information could improve
the handling of the extreme data values. It is suggested that this approach should be pursued in
future work. In contrast to kriging variance, the entropy depends not only on the distance to the
next data point, but also on the homogenity of the information. The calculated realizations can
also be used for design and numerical simulations, for example, to calculate failure probabilities
using stochastic or random finite element methods, as we are planning to do in future research.

CASE STUDY 2: FACIES MODEL FOR A METRO CONSTRUCTION SITE IN


MUNICH

A workflow for the geologic situation in Munich was created, based on the data collected
during the subsoil investigation campaign.
The subsurface in the Munich region is composed of Neogene and Quaternary formations
made up of fine- to coarse-grained sediments. In the Neogene, erosional debris from the Alps
was deposited in the Molasse Basin from Lake Constance to the Vienna Basin. The highest
member, the Upper Freshwater Molasse, was deposited in a lacustrine and limnic environment.
Eroded material from the Variscan Bohemian Massif is intermixed and contributes to the typical
mica content of the sands. Two main facies of these Neogene deposits can be distinguished: (1)
fluvially deposited sands and (2) clays deposited in still water. During the glaciation of the Alps
in the Quaternary Ice Age, over 100 meters of the Molasse sediments were eroded by meltwater
and large amounts of gravel were transported into the periglacial region of Munich. This led to
the formation of the so-called Munich (3) gravel plain.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 130

Figure 4 shows the workflow used to construct the random models. The coexistent
sedimentation of sands and clays have no clear geological body geometry, therefore indicator-
based interpolation and simulation methods are appropriate (Pyrcz and Deutsch 2014). This
indicator scheme is not used for the Quaternary gravel. SISIM is unsuitable to simulate interfaces
such as faults or erosion discordances, because the SISIM algorithm generates lenses of, or
example, the older layer within the younger one (Deutsch and Journel 1998). Therefore, erosion
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

is better represented as a layer whose interface can be uniquely identified. A 2.5D SGSIM of the
erosion surface was calculated as with the SGSIM method described in the first case study,
except that the variable is not the tip resistance but rather the height of the surface identified in
each borehole. Each simulation is transformed to 3D by classifying the grid above and below the
erosion surface.

Figure 4: Workflow and combination of Sequential Gaussian and Indicator Simulations.

The GSLIB algorithm SISIM_LM (see Deutsch and Journel 1998) was used to perform the
Indicator Simulation for tertiary sand and clay in order to quantify the uncertainty and determine
the most probable model. The SISIM_LM algorithm allows the incorporation of a spatially
variable mean value as Prior Information. Similar to the Ordinary Kriging method this mean
value was calculated from all boreholes in the model domain over the vertical axis. The mean is
constant in the horizontal direction and varies in the vertical direction.
Geological and subsoil investigations have been carried out at the emplacement site of a
future subway station in Munich. Twenty-three borehole logs were used to calculate a
probabilistic 3D model. One hundred SGSIM and SISIM_LM simulations were calculated on a
grid with a constant cell size of 10 m x 10 m x 0.5 m.
Figure 5 shows two 3D visualizations of the model, the boreholes and the bored pile wall.
The visualizations were generated with the program ParaView. The left figure shows the most
probable soil type and the right figure the uncertainty displayed as entropy for the clay layers. At
the location and near the boreholes, entropy is the lowest (0 or close to 0) and corresponds to low
uncertainty of model prediction.
The model will be used to improve the planning of the bored pile wall, which is used to
construct an excavation pit surrounded by a bored pile wall. The model enables the detection of
gaps in the confining clay layer and the evaluation of the risk of water inflow into the excavation.
It also assists with the planning of any additional investigations, quantity surveys, probabilistic
static design checks and optimization of the design.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 131
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5: Final model a) of the most probable facies; b) of the entropy of the predicted clay
layers.

CONCLUSION AND OUTLOOK

3D spatial uncertainty quantification is not common in geotechnical engineering and


traditional practice due to among others expensive or complex geoscience software, lack of
training and missing detailed guidelines in design codes. Information on the uncertainty of the
prediction regarding the subsoil structure and its properties is as important for the planning,
design and execution of construction projects as the prediction itself. The typical representation
and interpolation of subsoil layers is not suitable for the quantification and representation of
uncertainties and as such limits the risk assessment analysis of every infrastructure project.
The digitization of ground investigation data makes it even more urgent to further develop
and implement implicit modeling methods to quantify and communicate uncertainties. Currently
popular software for geotechnical subsoil modeling, such as Bentleys Leapfrog or others (see
Stütz and Herten 2020), do not provide simulation tools such as SGSIM and SISIM needed to
quantify uncertainty. Uncertainty quantification is a standard procedure in the oil and gas
industries and would help to mitigate risk in infrastructure projects such as tunnels, large
excavations, slope stability analysis, etc. In this paper, methods of statistical and geostatistical
simulation were illustrated in two case studies using CPT’s and borehole logs. Several measures
are available for evaluating different aspects of spatial uncertainty, e.g., kriging variance,
simulation variance, and entropy. The kriging variance depends only on the distance to the next
sample point. Geostatistical simulation can be used to determine likely soil model realizations
that can be used, for example, in probabilistic limit state analyses using the random finite
element method. The average of a high number of random realizations approximates the
interpolation of simple kriging. The simulated realizations of the ground can be used to spatially
compute the simulation variance, the probability of a soil type or facies, and the entropy, which
is a measure of the quality of the prediction.
In the future, it is intended to utilize further case studies to validate the suitability of the
herein proposed methods for quantifying uncertainties in subsoil modeling. The testing and
selection of suitable validation methods is planned, since they are a prerequisite for the
verification of the method presented above. In cross-validation, for example, parts of the data set

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 132

are removed creating a ‘test set’, and the simulation repeated with the remaining data. The new
prediction is then checked against the ‘test set’. Furthermore, different modeling or simulation
methods will be compared in terms of forecasting ability and suitability for quantifying
uncertainties. The methodology for quantifying the uncertainty of soil properties and fusion with
the 3D building ground model is the subject of our current research. Flexible analysis and
visualization tools for practical use will also be developed based on open-source standards and
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

programs.

REFERENCES

Ching, J., and Phoon, K.-K. (2020). “Constructing a Site-Specific Multivariate Probability
Distribution Using Sparse, Incomplete, and Spatially Variable (MUSIC-X) Data.” J. Eng.
Mech., 3, 1-15.
Deutsch, C., and Journel, A. (1998). GSLIB. Geostatistical Software Library and User’s Guide.
Second edition. Oxford University Press, Oxford, New York.
Gangrade, R. M., Grasmick, J. G., and Mooney, M. A. (2022). Probabilistic assessment of void
risk and grouting volume for tunneling applications. Rock Mechanics and Rock Engineering,
55(5), 2771-2786.
GeostatsPy. (2021). Geostatistical Library in Python. 1.0.0.
<https://github.com/GeostatsGuy/GeostatsPy> (May 10, 2022).
Grasmick, J. G. (2019). Modeling spatial geotechnical parameters uncertainty and quantitative
tunneling risks. Golden, CO: Colorado School of Mines.
Griffiths, D. V., Huang, J., and Fenton, G. A. (2009). “Influence of spatial variability on slope
reliability using 2-D random fields.” Journal of Geotechnical and Geoenvironmental
Engineering, 135(10), 1367-1378.
Kring, K., and Chatterjee, S. (2020). “Uncertainty quantification of structural and geotechnical
parameter by geostatistical simulations applied to a stability analysis case study with limited
exploration data.” International Journal of Rock Mechanics and Mining Sciences 125: 1-11.
Lloret-Cabot, M., Hicks, M. A., and van den Eijnden, A. P. (2012). “Investigation of the
reduction in uncertainty due to soil variability when conditioning a random field using
Kriging.” Géotechnique letters, 2(3), 123-127.
Pyrcz, M., and Deutsch, C. (2014). Geostatistical Reservoir Modeling. Second edition. Oxford
University Press, New York.
Robertson, P. K. (2010). “Soil behaviour type from the CPT: An update”. Proc. of the 2nd
International Symposium on Cone Penetration Testing, Huntington Beach, CA.
Shannon, C. (1948). “A Mathematical Theory of Communication.” Bell System Technical
Journal 3: 379–423.
Stütz, D., and Herten, M. (2020). “Evaluation von Software zur Generierung von
Baugrundschichtenmodellen.“ Geotechnik, 4: 275-282.
Wang, H., Wang, X., and Lian, R. (2020). Study of AI Based Methods for Characterization of
Geotechnical Site Investigation Data. Federal Highway Administration, Ohio.
Yang, R., Huang, J., Griffiths, D. V., and Sheng, D. (2017). “Probabilistic stability analysis of
slopes by conditional random fields.” Geo-Risk 2017, 450-459.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 133

Using MASW and Soil Borings to Characterize Heterogeneity of an Existing Solid Waste
Landfill

Kevin Foye, Ph.D., P.E., M.ASCE1


1
CTI and Associates, Inc., Farmington Hills, MI. Email: kfoye@cticompanies.com
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ABSTRACT

The variability of solid waste properties is typically greater than observed for natural soils
over comparable horizontal and vertical distances. Quantifying this variability is critical to the
success of solid waste landfill retrofit and final cover designs. This paper presents data from a
combined multichannel analysis of surface waves (MASW) and rotary soil boring investigation
of a 60-year-old, 8-hectare solid waste landfill. This investigation provided input to a final cover
design project. A comparison of the data illustrates where these technologies complemented each
other for characterization of subgrade properties and where they conflicted. In general, the
combined use of these technologies yielded valuable insights into subsurface conditions. The
data also yielded parameters useful for modeling the spatial variability of old solid waste that can
be used for future design analyses.

INTRODUCTION

The slope of final cover systems must be engineered to ensure drainage and cover integrity
throughout the design life of the facility, especially under the action of settlement. Accordingly,
the estimation of final cover settlement is an important engineering analysis activity in assessing
potential designs. Differential settlement, not total settlement is the most pertinent to this
problem. In particular, the creation of local depressions in landfill final cover systems can
prevent them from draining properly and minimizing infiltration through the liner system (Qian
et al. 2002).
Settlement of final cover systems is due to compression of both the landfill foundation and
the landfill waste mass. This compression is a function of the compressibility of the materials
and the loads imposed on them. Differential settlements therefore arise from both differences in
loading and differences in compressibility. Engineers can assess differences in loading directly
through consideration of the design geometry. In contrast, a complete assessment of differences
in compressibility is impossible due to a nearly limitless number of differences in material
composition, stress history, density, etc. that exist in a landfill. This problem is exacerbated for
waste materials, which generally exhibit greater heterogeneity than soils.
Foye and Soong (2010) introduced a methodology to simulate the differential settlement of
landfill final covers using random fields in conjunction with a simple isolated column settlement
model. In this methodology, random fields are generated to simulate the variation of subgrade
compressibility and subsequently compute the settlement at discrete points throughout the plan
area of the final cover. This process is repeated for multiple realizations of the random field in
order to generate a large population of post-settlement slopes.
A key feature of this probabilistic differential settlement analysis is the ability to explicitly
account for possible variations in the mechanical properties of the subgrade profile across the
domain of the settlement analysis. Of particular interest is the variation in compressibility across

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 134

the design profile. This variation results from differences in waste composition, density, and
moisture content. Shear wave velocity (vs) also depends on these properties. Therefore, MASW
surveys are a useful tool to quantify spatial variability in compressibility since variations in vs
can serve as a proxy for variations in compressibility. Because of the thousands of data points
collected by a MASW survey, this method provides sufficient data resolution to allow an
evaluation of spatial variability.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

In the following sections, a summary of a MASW survey of a 60-year-old, 8-hectare inactive


solid waste landfill is presented. In addition to quantifying subgrade variability directly using the
MASW vs profiles, inferences about the variation in waste composition and consistency across
the site could be made by comparing the MASW vs profiles to samples obtained from borings
into the landfill.
The project team completed 15 MASW survey transects and 10 intrusive borings at the
project site. Figure 1 presents the relative locations of the MASW transects and borings on an
aerial view of the landfill. The MASW survey included 15 transects arranged in a grid.
Following data reduction, model inversion, and post-processing, the project team evaluated the
resulting shear wave velocity profiles to delineate zones of interest for intrusive investigation.
The team selected boring locations such that each of the major trends in the profile was targeted
with at least one boring. The following sections describe both the comparison of these data sets
and the quantification of subgrade variability using MASW.

MASW SURVEY

The project geophysicist performed a total of 1,541 MASW shots using twenty-four 4.5 Hz
geophones connected to a Geometrics 24-Channel Geode seismograph. The geophones and cable
were attached to a “landstreamer,” which allows field personnel to rapidly move the entire
geophone array at one time during the survey testing. Based on the target depth range, each
MASW transect consisted of one spread of 24 geophones spaced 1.5 m apart and a total array
length of 35 m. This spacing resulted in a depth penetration range of 12-24 m below the ground
surface, depending on subsurface conditions and background noise at each transect location. The
energy source consisted of a 9-kg sledgehammer placed 3 m from the first geophone. For each
transect, a single “shot” was recorded, providing a 1D profile of shear wave velocity at the center
of the geophone array. The entire array was then shifted 1.5 m in line with the transect, and
another shot was performed. This procedure continued until the end of the array reached the
limits of the survey area.
The project geophysicist used Geometrics Seismodule Controller acquisition software
connected via a field laptop to the seismograph to monitor noise and observe/record each shot
gathered during the MASW survey. Subsequent to field data collection, the geophysicist
processed all seismic files grouped into each of the seventeen transects (two of the 15 transects
were divided as shown in Figure 1) using the Kansas Geological Survey Surfseis 6.0 data
processing software. This software generates a dispersion curve image for each individual
seismic shot. The dispersion curve images are analyzed to identify the fundamental mode
relationship between velocity and frequency, and the dispersion curve within the highest
amplitude zone of the velocity/frequency relationship is selected by the geophysicist. When a
curve has been chosen for all seismic shots in a transect, the software completes a dispersion
curve inversion process to generate a 1D profile of shear wave velocity with depth at each shot
location. The geophysicist assumed 10 layers for this inversion. The implications of this choice

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 135

are mentioned in the discussion. The geophysicist combined all 1D profiles for each transect into
a 2D graphical cross section of vs using a kriging process. Following this process, the
geophysicist generated 2D cross sections of vs for each of the seventeen transects. Figure 2
presents the resulting graphical vs profile for Transect 4. A graphical representation of Boring 2
is also plotted against the vs profile to aid interpretation.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Map of Project Site Showing MASW Transect and Boring Locations. Transects 8
and 9 were divided into two parts each to accommodate rough terrain.

Figure 2. Interpreted Shear Wave Velocity Profile Through Transect 4.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 136

INTRUSIVE BORINGS

Following review of the MASW profiles and selection of representative shear wave velocity
zones, the project team performed drilling, sampling, and description of in-place landfill
materials as follows. Borings were advanced using a hollow-stem auger drill rig mounted on a
tracked all-terrain vehicle capable of traversing the site. Split spoon samples were collected using
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Standard Penetration test (SPT) methods to directly observe materials in-place at each boring
location. SPTs were be performed at 0.7-m depth intervals. Documentation of the drilling
activity included boring logs with the following information: visual description of all samples
(color, consistency, texture), Standard Penetration Test (SPT) blow counts, and notable voids or
changes in profile.
The intrusive borings encountered the following general subsurface materials (strata with
waste and waste mixed with soil were encountered at various depth intervals throughout the site):
• Cover soils: silty clay fill, silty sand fill, sand fill, and sand and gravel fill to depths
ranging from 0.3 to 1.3 m below surface elevation. The encountered materials consisted
largely of the following primary components: Silty clay fill, Silty sand fill, Waste, Sandy
silt fill, Wood fill, Sand and gravel fill, Silty, clayey sand fill, Gravelly clay fill, Brick
fragments, and Lean clay fill. Varying amounts of secondary components observed
included sand, clay, silt, gravel, waste, wood, and brick.
• Fill Soil and Waste: In places where the soil content of encountered landfill debris
dominates the material behavior (i.e., waste included in a primarily soil matrix), the
descriptor “Fill Soil and Waste” is used.
• Waste: Solid waste was encountered with varying degrees of degradation and soil
content. The individual materials comprising the waste were indiscernible in
approximately 40 percent of samples. Discernable waste materials included wood, plastic
film, plastic, construction and demolition debris, brick fragments, fabric, bone, and
rubber.
• Native Soils: Several borings extended to the natural clays underlying the site. The
encountered native soils consisted of predominantly stiff to very stiff lean clay with
generally trace sand or sandy silt.

DISCUSSION OF RESULTS

To aid MASW profile interpretation, the geophysicist generalized the detailed strata
documented in the boring into the following categories: Surficial soils; Soil containing waste
(“soil & waste”); Predominantly waste with little soil (“waste”); deeper native soil (lean clay and
sandy silt). These categories are reflected in the generalized profile shown in Figure 2. Table 1
presents a summary of a comparison of the MASW and intrusive boring observations.
In general, the north portion of the landfill (north of MASW Transect 7) exhibits lower vs
ranges within the waste deposit, and a more consistent shear wave velocity transition with
increased depth from the ground surface into the native soil underlying the waste. In contrast, the
southern portion of the landfill (south of MASW Transect 7) exhibits increased variability in vs
within the waste deposit and an overall greater velocity range. Review of the boring logs
suggests these differences may be, in part, the result of greater percentages of waste, decreased
soil content, increased saturation, and/or changes in the overall type of waste that is present in
the southern portion of the landfill.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 137

The horizon between waste and native soil at depth is well-represented in the MASW vs
profiles. The profiles show a consistent transition to higher vs between elevations of 180 – 186 m
among all of the transects performed at the site, corresponding to the observations made at
nearby borings.

Table 1. Summary of MASW and Soil Boring Profile Comparisons


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Foundation Soil
Profile Notes Elevation (m)
typical vs
transect MASW boring (m/s) MASW boring
alternating high and low vs with alternating soil and soil &
1 90-270 183 n/a
depth waste
alternating high and low vs with alternating soil, waste, and
2 90-300 180 n/a
depth soil & waste
consistent transition low to high alternating waste and soil &
3 90-180 181 183
vs waste
consistent transition low to high shallow boring into soil &
4 70-180 181 181
vs waste
inversion produced a high-to- alternating waste and soil &
5 70-180 183 184
low vs sequence near boring waste
consistent transition low to high
6 fill soil & waste over waste 70-180 183 183
vs
consistent transition low to high alternating soil, waste, and
7 70-230 183 183
vs soil & waste

fill soil near surface, then


8A, 8B zone of higher vs near surface alternating fill soil & waste 90-300 180 n/a
and waste
significant vs variability near
9A, 9B n/a 70-300 186 n/a
surface
consistent transition low to high primarily soil & waste,
10 70-240 n/a n/a
vs, some pockets of higher vs some pockets of soil
highly variable vs in north, less
11 n/a 70-260 181 n/a
variable vs in south
consistent transition low to high
12 n/a 60-170 181 n/a
vs
consistent transition low to high
13 n/a 70-230 183 n/a
vs
consistent transition low to high, primarily soil & waste,
14 70-230 180 184
some pockets of higher vs some pockets of waste
alternating waste and soil &
15 highly variable vs 100-330 183 183
waste

EVALUATION OF SPATIAL VARIABILITY

The probabilistic differential settlement analysis described in the introduction uses a random
field generator to assign different possible compressibility values to the subgrade according to

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 138

key statistics that describe the variability of compression index. The first parameter, standard
deviation, describes the overall variability in compressibility possible over the site. The second
parameter, , sometimes termed “scale of fluctuation” describes the rate at which compressibility
varies in space according to the following decay function:

|𝜏|
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

𝜌(𝜏) = exp (− 𝜃 ) (1)

where  is the correlation coefficient and  is the lag distance. Values of  equal to 1 indicate
perfect correlation, which generally only occurs when comparing a variable to itself for  = 0
(i.e., a measurement taken in exactly the same place). As lag distance  increases (i.e.,
measurements taken further apart from each other)  decreases, indicating less correlation
between the measurements. Values of  appropriate for use in probabilistic analyses are difficult
to obtain from typically sparse subsurface data. Accordingly, authors modeling differential
settlement have relied on conservative assumptions of , such that estimates of differential
settlement are greater than expected (see for example, Fenton et al. 1996). Data that can provide
additional insight into the selection of  are therefore highly desirable for analysis and design
using probabilistic methods.
To evaluate an appropriate value of  to analyze differential settlement at the project site, 2D
correlograms of the vs profiles were first generated using Surfer, a spatial data analysis tool.
Correlograms plot correlation coefficient  versus lag distance . An example 2D correlogram is
shown in Figure 3. The colors in Figure 3 correspond to values of .  = 1 at the origin of the
plot (marked “0,0” in Figure 3) and decreases with increasing lag distance from the origin. Cross
sections through the 2D correlograms were extracted in both the vertical (‘y’) and horizontal
(‘x’) direction as depicted in Figure 3.

Figure 3. Example Two-Dimensional Correlogram

The extracted correlogram cross sections were compared to the values generated using Eq. 1.
For each correlogram cross section the project team optimized  to minimize the error between
the fitted function and the data. The resulting values of fitted  parameters are presented in Table
2. Example fitted curves are also presented for Transect 4 in Figures 4 (‘x’ direction) and 5 (‘y’
direction).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 139

Table 2 shows that  in the horizontal direction is about 10 times  in the vertical direction,
which is consistent with the expected pattern of waste lift placement, compaction, and
consolidation for solid waste landfills. However, it is also likely that values of  in the vertical
direction are a direct consequence of the MASW profile inversion performed. The software was
used to find a trial shear wave velocity profile with 10 layers that produces a similar theoretical
dispersion curve that matches the experimental dispersion data extracted from the geophone
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

measurements in the MASW survey. With the total profile thickness modeled being typically 18
m, a vertical  = 1.5 m corresponds closely with the 1.8 m layer thickness used to invert the
MASW data.

Figure 4. Transect 4 Correlogram – ‘x’ Direction

Figure 5. Transect 4 Correlogram – ‘y’ direction

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 140

In contrast, the horizontal  values are primarily affected by variations between individual
MASW shots, which are less subject to interpretation during data reduction and inversion.
Because any given MASW shot represents a spatial average of all material beneath the geophone
array, horizontal variation in these profiles is indicative of variation in a “moving average” of
properties across the profile. The typical value of horizontal  obtained (27 m) is much greater
than the MASW shot spacing of 1.5 m, suggesting that significant fluctuations in in-place waste
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

properties in the horizontal direction occur over larger distances, as expected. The choice of
horizontal value of  affects the estimation of differential settlement between different points on
the final cover system. Therefore, this parameter is the more significant result of the data
analysis.

Table 2. Scale of Fluctuation () Calculated from MASW Shear Wave Velocity Profiles

Transect Horizontal direction  (m) Vertical direction  (m)


1 45.6 2.4
2 27.1 1.3
3 31.4 1.6
4 23.3 2.0
5 17.1 1.8
6 19.2 1.8
7 36.9 2.6
8A 26.7 2.7
8B 17.6 1.8
9A 11.2 2.7
9B 24.8 2.5
10 12.3 1.5
11 31.3 2.5
12 55.5 2.4
13 50.9 1.7
14 64.3 2.2
15 53.0 3.2
median 27.1 2.2

CONCLUSIONS

The following conclusions can be drawn from the above data and discussion:
• The MASW profiles generated for this project allowed a qualitative assessment of
variations in landfill properties. However, the ability of these profiles to reliably
discretize particular zones of material is limited.
• The MASW profiles and soil boring logs show a sequence of alternating soil and waste
layers within the landfill footprint. Geostatistics performed on the MASW data confirm a
layered structure, with mechanical properties varying about 10 times more rapidly in the
vertical direction than in the horizontal direction. This finding is consistent with
landfilling practices. However, it may also be a consequence of the MASW data
inversion techniques used for the project.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 141

• Estimates of horizontal  from MASW profiles appear to be appropriate, especially since


the horizontal variation of inverted profiles is directly related to the measured MASW
data.
• Interpretation of the bottom of waste (i.e., the top of clay) from the MASW vs profiles
correlated well with the physical samples obtained from intrusive borings. Accordingly,
this technique may be applicable to future investigations, provided a limited number of
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

intrusive borings are similarly used to aid profile interpretation.


• Additional insights into the MASW data inversion process for old landfills in general and
the project site in particular are possible through a deeper comparison of the boring and
MASW results.

ACKNOWLEDGEMENTS

The MASW survey and data reduction was performed by Pyramid Geophysics, Greensboro,
NC. The author is grateful to Mr. Eric Cross, P.G. for his work on this project.

REFERENCES

Baecher, G. B., and Christian, J. T. (2003). Reliability and Statistics in Geotechnical


Engineering. Wiley, Hoboken, NJ, 605 p.
Fenton, G. A., Paice, G. M., and Griffiths, D. V. (1996). “Probabilistic Analysis of Foundation
Settlement.” Uncertainty in the Geologic Environment: from Theory to Practice, GSP No.
58, ASCE, Reston,VA: 651-665.
Foye, K. C., and Soong, T. (2010). “Simulating Differential Settlement of Landfill Foundations
using Random Fields” GeoFlorida 2010. West Palm Beach, Florida.
Qian, X., Koerner, R. M., and Gray, D. H. (2001). Geotechnical Aspects of Landfill Design and
Construction. Prentice Hall, Upper Saddle River, NJ, 717 p.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 142

Using Machine Learning to Predict Shear Wave Velocity

Longde Jin, Ph.D., P.E.1; Andrew Fuggle, Ph.D.2; Haley Roberts3;


and Christian P. Armstrong, Ph.D., P.E.4
1
WSP Golder, Atlanta, GA. Email: longde.jin@wsp.com
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
WSP Golder, Atlanta, GA. Email: andrew.fuggle@wsp.com
3
Georgetown Univ., Washington, DC. Email: hne7@georgetown.edu
4
WSP Golder, Denver, CO. Email: christian.armstrong@wsp.com

ABSTRACT

Cone penetration testing (CPT) is widely used for geotechnical characterization of subsurface
materials. Shear wave velocity (Vs) measurements provide very useful information for
engineering purposes and can usually be added to land-based CPT testing programs for a
relatively minor increase in cost and time. However, performing Vs measurements from a
floating platform on submerged materials such as impounded tailings and coal ash is a challenge
due to the lack of available equipment required to generate shear waves from the mudline. To
solve this problem, the authors developed a new model using machine learning techniques to
estimate Vs based on conventional CPT measurements without direct Vs measurements. The
model uses a random forest algorithm implemented in Python for solving this regression
problem. The new model is shown to provide more accurate estimates of Vs than methods
widely used currently. The technique also provides for the possibility of generating continuous
Vs profiles with greater vertical resolution than currently obtained using discrete measurements.

INTRODUCTION

A widely applied approach for estimating Vs from CPT data is through empirical correlations
(Mayne 2006, Andrus 2007, Robertson 2009, and Wair et al. 2012). Applying this approach to
impounded tailings and coal ash is problematic because few such case histories were used in the
development of these methods. Another approach is to use bender element testing of undisturbed
samples in a laboratory. This approach is also problematic due to the difficulty in obtaining high-
quality, undisturbed samples of very loose saturated materials and the effects of an unknown
degree of disturbance from sampling and transport on measured properties. Due to these
concerns, in-situ testing, especially Seismic Cone Penetration Testing (SCPT), are often used as
the primary means for direct measurement of shear wave velocity.
The in-situ data compiled for this study consists of 431 Vs measurements from 21 SCPT
soundings. 13 SCPTs are from four coal combustion residual (CCR) impoundments in the
Eastern United States and eight SCPTs are from one tailings storage facility.
The CCR data included in this study is from areas of CCR impoundments where fly ash is
predominant. Fly Ash is typically a non-plastic silt to sandy silt with a sand content of less than
35%. Specific gravity of Fly Ash varied from 1.93 to 2.46 with an average value of 2.19 (Fuggle
et al. 2022). Dry unit weight ranged from 3.5 to 17.2 kN/m3 with an average dry unit weight of
10 kN/m3 .
The tailings studied for this work, consists of molybdenum tailings which is classified as a
silty sand or a sandy silt. Fines content ranges from 22% to 84%, with an average of 45% (Been
et al. 2012). A unit weight of 15 kN/m3 was used for estimating the vertical stress.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 143

Statistical analysis of the raw SCPT measurements is presented in the following section.
Three different data sets are then used for further analysis. The first data set includes only Fly
Ash data, the second data set includes only the tailings data, and the third data set is a
combination of Fly Ash and tailings data. Comparisons to widely used empirical correlations are
also presented.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

SEISMIC CONE PENETRATION TESTING (SCPT)

SCPT typically records base data (tip stress [qt], friction sleeve [fs], and penetration
porewater pressure [u2]) every 2 cm. Vs measurements are typically made at intervals of 1 or 2
meters, or even greater. The base CPT measurements are averaged over the depth range between
each Vs measurement. Statistics of the CPT base data for the combined dataset, estimated total
and effective vertical stress, 𝜎𝑣 and 𝜎𝑣′ , hydrostatic pressure, u0, are summarized in Table 1 and
presented in Figure 1, with the Fly Ash, tailings, and combined data sets shown.

Table 1. SCPT combined dataset statistics

Range Statistics
# oftestsa
Property Estimation Method Min Max 𝜇 𝜎 Median QCDb
(# of borehole)
Depth (m) CPT Base Data 431b 2 126 29 27 18 0.61
(21)
qt CPT Base Data 0.10 31.64 8.37 8.78 3.99 0.87
(MPa)
fs CPT Base Data 0.00 0.95 0.12 0.16 0.05 0.78
(MPa)
u2 CPT Base Data -0.02 0.47 0.09 0.11 0.02 0.98
(MPa)
𝜎𝑣𝑜 Estimated using 0.03 1.89 0.44 0.40 0.29 0.59
(MPa) typical
unit weight for the
material

𝜎𝑣𝑜 𝜎𝑣𝑜 − 𝑢0 0.02 1.28 0.38 0.36 0.19 0.70
(MPa)

u0 Estimated from 0.00 0.61 0.06 0.10 0.01 1.00


(MPa) CPT

Ic Robertson and 1.77 5.55 2.77 0.52 2.65 0.13


Wride (1990)

Note: a. All data is averaged in the depth window of each shear wave velocity measurement
b. Quartile Coefficient of Dispersion (QCD) is defined as (Q3-Q1)/(Q3+Q1) where Q1 and Q3
are first and third quartiles for each data

The Fly Ash data shows a typical response of low qt, low fs, and high u2. Tip stress ranges
from 0.1 to 6.2 MPa with an averaged value of 1.1 MPa; friction sleeve ranges from 0.00 to 0.20
MPa with an averaged value of 0.03 MPa; and penetration pore pressure ranges from 0.0 MPa to
0.5 MPa with averaged value of 0.2 MPa. All the Fly Ash data was obtained from less than 30
meters below the mudline.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 144
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. SCPT dataset – input parameters

In comparison, the tailings dataset shows a wider range response of qt, fs, and nearly no pore
pressure response. Tip stress ranges from 0.5 to 31.6 MPa with an averaged value of 13.8 MPa;
and friction sleeve ranges from 0.00 to 0.95 MPa with an averaged value of 0.19 MPa. The
maximum depth of the tailings dataset is 126 meters below ground surface.
The soil behavior index, Ic (Robertson and Wride, 1990), indicates that 63% of the Fly Ash
data have clay-like behavior and 62% of the tailings data have sand-like behavior. The combined
dataset shows a wider range of behavior, i.e., 37%, 23%, 28%, of sand mixtures-, silt mixtures-,
and clay-behavior, respectively. Frequency of the Ic for each dataset is summarized in Table 2.
The field-measured Vs of the Fly Ash dataset ranges from 48 to 267 m/s, and the measured
shear wave velocity of the tailings dataset ranges from 88 to 455 m/s. The distribution for each
dataset is shown in Figure 2. The Fly Ash Vs measurements are generally less than 200 m/s
while the tailings Vs measurements have a wider range. Table 3 summarizes the statistics of the
measured Vs from SCPTs. It also shows the range of values for three widely used CPT based
empirical correlations applied to the combined dataset.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 145

Table 2. Frequency (%) of soil behavior index of dataset

Soil Behavior
Ic Fly Ash Tailings Combined
Type
Gravelly Sand < 1.31 0 0 0
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Sands 1.31 – 2.05 0 13 7


Sand Mixtures 2.05 – 2.60 4 62 37
Silt Mixtures 2.60 – 2.95 21 23 23
Clays 2.95 – 3.60 63 2 28
Organic Soils > 3.60 11 0 5

Table 3. Measured and estimated vs(m/s)

Range Statistics

# of testsa
Dataset Method (# of Min Max 𝜇 𝜎 Median QCDb
borehole)
Fly Ash SCPT Measured 184 48 267 143 41 144 0.17
(13)

Tailings SCPT Measured 247 88 455 249 82 251 0.25


(8)

Combined SCPT Measured 431 48 455 204 86 183 0.31


(21)

Combined SCPT Mayne (2006) 431 60 504 239 97 221 0.32


(21)

Combined SCPT Andrus (2007) 431 62 635 260 152 189 0.47
(21)

Combined SCPT Robertson (2009) 431 38 652 261 155 193 0.48
(21)

Note: a. All data is averaged in the depth window of each shear wave velocity measurement
b. Quartile Coefficient of Dispersion (QCD) is defined as (Q3-Q1)/(Q3+Q1) where Q1 and Q3
are first and third quartiles for each data

MACHINE LEARNING APPROACH TO Vs ESTIMATION

While there are many machine learning (ML) algorithms and methods, this study selected the
Random Forest algorithm because of its ease of use and performance. The tree-based nature of
the algorithm allows for the results to be more easily interpreted and communicated compared to
more complicated algorithms, such as neural networks. Additionally, the Random Forest
algorithm provides ease of tuning as compared with other tree-based models such as XGBoost
(Chen et al. 2015). Considering this, a Random Forest model is used for this study.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 146
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. SCPT dataset – field measured Vs

The Random Forest algorithm is a ML method for predicting features of new data based on
previous data using either regression or classification methods built via computer programming
(Pedregosa et al. 2011). Models are trained on existing data and applied to new data to predict
features and provide insights. The building block of the Random Forest algorithm is the Decision
Tree which is a prediction algorithm that successively splits training data based on its features
and then predicts new data based on these "decision" splits. The Random Forest model uses a
collection of Decision Trees (each with different "splits") built based on the training data to
predict the testing data’s label (classification) or value (regression) based on majority voting
from the many trees in the forest.
In this study, the input data is base data from SCPTs along with some variables that are
independent of the base CPT data, such as effective stress (𝜎𝑣′ ) and hydrostatic pore pressure
(u0). The input data is randomly split into two data sets: 80% to be used as training data to
construct the Random Forest model, and 20% to be used as testing data for the built model to
predict features of and be evaluated for accuracy. The approach to training and testing data sets
is presented in Figure 3.

Figure 3. Approach to training and testing data sets

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 147

The Random Forest model’s hyperparameters can be adjusted, or tuned, to optimize


prediction performance (Pedregosa et al. 2011). In this study, five hyperparameters are adjusted:
n_estimators, min_samples_split, min_samples_leaf, max_depth, and bootstrap. A tuning grid is
built to search through a specified list of integers for each tuned parameter and identify the
optimal combination of these parameters for higher accuracy. This grid consists of n_estimators
(every 100 units from 100 to 1700), min_samples_split (every 2 units from 2 to 9),
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

min_samples_leaf (every 1 unit from 1 to 6), max_depth (every 20 units from 10 to 110). All
other hyperparameters are set as their default values. A randomized search algorithm with 5-fold
cross-validation (to reduce overfitting) is used to test 100 different combinations of the specified
parameter values. The parameters producing the minimal error are returned and used for the final
Random Forest model. It should be noted that using a tuning grid with a greater number of
combinations of parameters may provide higher model accuracy, at the cost of computational
time and power. However, for the purposes of this study, the hyperparameter tuning process
performed is regarded as sufficient.

MODEL PERFORMANCE

This section examines how well the ML can predict Vs from base CPT data. Figure 4a shows
that the ML model trained only on Fly Ash data does not predict tailings Vs well, especially
above about 200 m/s. This is expected as the Fly Ash dataset does not contain many data points
with higher Vs. The model trained on Fly Ash data shows much better performance in predicting
previously unseen (to the model) Fly Ash data represented by the green squares. Table 4
indicates that the mean absolute error (MAE) is 19.5 m/s when applying the Fly Ash-trained ML
model to Fly Ash but shows a MAE of 85.6 m/s when applying that same model to tailings.
Figure 4b shows that the ML model trained only on tailings data predicts the unseen tailings
data well (purple squares) but does not accurately predict the Fly Ash data. Table 4 indicates that
the MAE is 20.1 m/s when applying the tailings-trained ML model to tailings data. Applying the
tailings-trained ML model to Fly Ash gives a MAE of 27.5 m/s. The tailings-trained ML model
is a better predictor of Vs in Fly Ash than the Fly Ash-trained ML model is of Vs in tailings.

(a) (b)

Figure 4. Vs comparison between measured and interpreted. (a) ml model trained only
on fly ash data, (b) ml model trained only on the tailings data.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 148

Table 4. Accuracy of predicted values of testing dataset

Trained on 80% of Trained on 80% of Trained on 80% of


Fly Ash Dataset Tailings Dataset Combined Dataset
20% Fly 100% Fly
Testing Set 100 % Tailings 20% Tailings 20% Combined Data
Ash Ash
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

MAE (m/s) 19.5 85.6 27.5 20.1 21.9

Relative Error
13.6% 34.4% 19.2% 8.1% 10.7%
(%)
Training Data
147 147 197 197 344
Points

The MAE for the ML model trained on the combined dataset is 21.9 m/s when applying the
model to a combined testing set of tailings and Fly Ash. This is a very slight degradation in
predictive accuracy from the ML models trained and tested specifically on one single material.
Figure 5 shows how a ML model trained on the combined data set performs when predicting
Vs in both Fly Ash and tailings. Also shown in Figure 5 is how existing correlations perform on
this same dataset.

Figure 5. Vs comparison between measured and interpreted. plots only show the testing set
from combined dataset

Table 5 shows a comparison between existing correlations and the ML model. The ML
model shows a MAE of 21.9 m/s which compares very favorably to the MAE of 46.3 m/s, 70.3
m/s, and 72.6 m/s for the Mayne, Andrus, and Robertson models, respectively. The existing

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 149

correlations did include Vs data over about the same range as the data from this study but did not
include the actual data used in this study as part of their datasets so it is expected that the ML
model would show greater predictive accuracy. However, the magnitude of the improvement
(reducing the MAE to between half and one-third) is, in our opinion, a particularly encouraging
result.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Table 5. Comparison of ML and existing models.

ML Model
Metric Mayne (2006) Andrus (2007) Robertson (2009)
(This study)
MAE (m/s) 21.9 46.3 70.3 72.6
Relative Error (%) 10.7% 23% 35% 36%
Training Data Points 344 161 185 1035

Figure 7 presents how the ML model trained on the combined dataset can be used to interpret
the continuous in-situ Vs profile. The first three rows show the ML model predictions for Vs in
Fly Ash and the lower two rows show the prediction of Vs in tailings. These results are
particularly encouraging although care must be used in about the upper 5m of any sounding.

Feature Importance

One benefit of the ML model is that the importance of different variables can be estimated.
The importance of a particular variables is estimated using the Gini coefficient. Results of this
analysis are shown in Figure 6. The Gini coefficient ranges from 0 to 1, where 1 stands for the
most important factor, and 0 stands for the least important factor.

Figure 6. Feature importance for different training set. (a) Model trained only using fly ash
data, (b) model trained only using tailings data, (c) model trained using combination data.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 150

Results show that for the Fly Ash dataset (Figure 6a), the effective stress, sleeve friction and
penetration porewater pressure are the most important factors when estimating Vs. In contrast,
for the tailings dataset (Figure 6b) the effective stress and tip stress are the most important
factors. For the combined dataset (Figure 6c) the tip stress and effective stress are estimated to be
the most important, with the hydrostatic and penetration porewater pressures having almost no
influence on the results.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 7. Vs comparison between measured and interpreted. The first three rows are the
prediction of Vs in Fly Ash, and the last two rows are the prediction of Vs in tailings.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 151

A general site-specific correlations form provided by Wair et al. (2012) suggests the shear
wave velocity can be estimated from 𝑞𝑡 , 𝑓𝑠 , 𝑎𝑛𝑑 𝜎𝑣′ . This correlation captures the three factors
estimated to be most important from the combined dataset using the ML approach.

CONCLUSION
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

This paper has presented a new model using machine learning techniques to estimate Vs
based on conventional CPT measurements without direct Vs measurements. The newly
developed estimator of Vs appears more accurate than existing approaches with the reduction in
MAE to about half to one-third of current approaches. This approach can be used for site- or
material-specific models, or for more broadly applicable models generalized to all material types.
The use of larger and more varied training datasets would be required to produce a more broadly
applicable model. The model estimated the importance of various input variables and identified
qt, fs, and 𝜎𝑣′ as the most significant inputs for the combined dataset. Interestingly, the material-
specific models showed a different ranking for the importance of the input variables. The use of
Ic, or some other indicator of material type, as an input variable may improve the model
predictions. The authors plan to explore additional ML methods to bolster the study. We are also
considering making a platform available for others to explore the approach.

ACKNOWLEDGEMENTS

The authors would like to thank WSP Golder and our valued clients for the opportunity to
publish this paper and to work on challenging and interesting projects. The authors would also
like to thank the reviewers who provided valuable comments and suggestions.

REFERENCES

Andrus, R. D., Mohanan, N. P., Piratheepan, P., Ellis, B. S., and Holzer, T. L. (2007). Predicting
shear-wave velocity from cone penetration resistance, Proc., 4th Inter. Conf. on Earthq.
Geotech. Eng., Thessaloniki, Greece.
Been, K., Crooks, J. H. A., and Jefferies, M. G. (1988). Interpretation of material state from the
CPT in sands and clays. Penetration Testing in the UK: Proc. of the Geotechnology
Conference, Thomas Telford Publishing, 215-218.
Chen, T., He, T., Benesty, M., Khotilovich, V., Tang, Y., Cho, H., and Chen, K. (2015).
Xgboost: extreme gradient boosting. R package version 0.4-2, 1(4), 1-4.
Fuggle, A., Jin, L., and Hebeler, G. (2022). Classification and Characterization of Ponded Coal
Combustion Residuals Using in Situ Methods. In Geo-Congress 2022 (pp. 525-535).
Mayne, P. W. (2006). In situ test calibrations for evaluating soil parameters. Proc.,
Characterization and Engineering Properties of Natural Soils II, Singapore.
Pedregosa, F., Varoquaux, G., Gramfort, A., Michel, V., Thirion, B., Grisel, O., and Duchesnay,
E. (2011). Scikit-learn: Machine learning in Python. the Journal of machine Learning
research, 12, 2825-2830.
Robertson, P. K. (1990). Soil classification using the cone penetration test, Canadian Geotech.
J., 27(1):151–158.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 152

Robertson, P. K. (2009). Interpretation of cone penetration tests – a unified approach, Canadian


Geotech. J., 46(11):1337–1355.
Wair, B. R., DeJong, J. T., and Shantz, T. (2012). Guidelines for estimation of shear wave
velocity profiles. Pacific Earthquake Engineering Research Center.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 153

Magnetic Resonance Imaging for Pore Water Mapping in Soils

Karam A. Jaradat, Ph.D., P.E., A.M.ASCE1; Maosen Wang, Ph.D.2;


and Sherif L. Abdelaziz, Ph.D., P.E., A.M.ASCE3
1
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

VA. Email: karam@vt.edu


2
Farlin Biomedical Research Institute, Virginia Tech Carilion, Roanoke, VA.
Email: maosen@vtc.vt.edu
3
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
VA. Email: saziz@vt.edu

ABSTRACT

This paper reviews the application of magnetic resonance imaging (MRI) for monitoring pore
water in different porous media. The physics of MRI and different imaging techniques are first
presented in detail, and studies focusing on quantifying water and water flow in porous media are
reviewed. Then, we present the results of MRI scans for a poorly graded sand (badger sand),
scanned utilizing a 9.4 Tesla MRI scanner. The specimen was saturated with deionized (DI)
water in a PVC container, then placed inside the magnetic resonance (MR) scanner for fast low
angle shot (FLASH) sequence imaging. The collected 2D images were then processed, and a 3D
tomography for the pore water distribution was reconstructed. The volumetric water content (wc)
for the badger sand specimen was then calculated by three-dimensionally segmenting of the pore
water component acquired using the MRI. The calculated water content agreed with the water
content used to prepare the specimens, validating the adopted scanning parameters and
methodology. Such results pave the way for more advanced analysis of pore water in soils under
steady state or varying water flow.

INTRODUCTION

The process of water flow controls various engineering phenomena in porous media such as
seepage, contaminants transport, recharging of aquifers, and surface and internal erosion.
Considering the heterogenous nature of soils and the interconnected pores with varying sizes,
water flow occurs at different scales in the micro- and the macro-pores. In most engineering
applications, water flow is simply modeled using a so-called hydraulic conductivity coefficient
that indicates on average how fast water flows in the pore structure. Geotechnical engineers often
use laboratory experiments (i.e., permeability tests) to determine how fast water flows through a
controlled (prepared) soil stratum under constant or falling head conditions (ASTM, 2016,
ASTM, 2019), and slug tests are used in the field to get the in-situ hydraulic conductivity of
natural soils (Zurbuchen et al. 2002). While these methods have been used successfully for
decades as a measure of water flow in soils, the direct observation of flow, whether in the lab or
the field, is difficult due to the opaque nature of these systems. Moreover, some processes (e.g.,
internal erosion) are characterized by the continuous detachment and removal of soil material
over time (Robbins and Griffiths, 2018), leading to an unpredicted sudden failure of the soil
mass. Therefore, a mean of monitoring water flow in porous media at different scales is needed
to understand the mechanism of soil failure, and to validate advanced particle-based numerical
models of fluid-soil interactions.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 154

Over the past two decades, advanced imaging and tomographic methods (e.g., FIB-SEM, and
x-ray computed tomography) have been widely used to develop images of soil structures
(Darbari et al., 2017, Taina et al., 2008). Magnetic resonance imaging (MRI) has been also used
to map water distribution in porous media (Amin et al., 1994, Greiner et al., 1997, Deurer et al.,
2002, Pohlmeier et al., 2010, Haber-Pohlmeier et al., 2010). As it directly probes the substance
of interest (i.e., water), MRI is suitable to study water flow processes in soils. Therefore, this
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

paper reviews the application of magnetic resonance imaging for monitoring water and water
flow in soils. Preliminary MRI scans a sandy specimen are also presented and recommendations
are given for future studies.

PRINCIPLES OF MAGNETIC RESONANCE IMAGING

Magnetic Resonance Imaging, MRI (also known as Nuclear Magnetic Resonance, NMR) is
an imaging tool that produces three-dimensional detailed mapping (i.e., imaging) of a targeted
nucleus (e.g., 1H), often used for disease detection, diagnosis, and treatment monitoring. The
highest sensitivity and best resolution for MRI is obtained with hydrogen (1H) as the detected
nuclear species (Pearl and Magaritz, 1993), and therefore it is suitable to study flow
characteristics of hydrogen-bearing molecules, such as water (H2O), in porous media.
The principles of MRI are schematically shown in Fig. 1. When water protons are initially
excited by a strong static magnetic field (B0) aligned along the axis of the MRI scanner (Z-
direction), they tend to align with it possessing a magnetization vector (M) that is parallel to B0
(Callaghan, 1991, Nestle et al., 2002, Lascialfari and Corti, 2007). Upon introducing a weaker
radiofrequency (RF) pulse (B1) and if it’s sufficiently long (90°), the protons get disrupted and
align in a different direction, causing M to tilt and water protons to store electromagnetic energy.
The magnetization vector, however, tends to relax and return to its initial state to restore
Boltzmann equilibrium; that is, parallel to the strong magnetic field B0.

Figure 1. Basic principles and physics of Magnetic Resonance Imaging (MRI).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 155
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Simplified diagram for spin-echo sequence, also termed multi-slice multi-echo
sequence, MSME (adapted from Haber-Pohlmeier et al., 2021).

Relaxation occurs in two schemes known as spin-lattice relaxation in the Z-direction


(longitudinal), and spin-spin relaxation in the XY plane (transverse). The rate of the spin-lattice
relaxation is described by T1 and the rate of the spin-spin relaxation is reflected by T2. The
transverse relaxation can be divided into at least two components, relaxation due to dephasing in
external magnetic field inhomogeneities, and relaxation due to random dephasing (Haber-
Pohlmeier et al., 2010). The former effect strongly dominates the overall relaxation but can be
eliminated by the creation of a spin echo, where the dephased magnetization vector in the
transverse direction is rephased by a 180° pulse (Haber-Pohlmeier et al., 2010). A period of
several milliseconds is required between excitation and detection, commonly known as the time-
to-echo (TE). The time between successive pulse sequences is referred to as repetition time (TR).
The process of excitation-relaxation causes the protons to release the stored electromagnetic
energy as a voltage signal, which is detected and processed by the MRI machine, as
schematically depicted in Fig. 2 for a multi-slice multi-echo sequence (MSME). Both T1 and T2
are function of the medium.
Imaging and signal intensity. One of the most common imaging protocols adopted in MRI
is the multi-slice, multi-echo sequence (MSME) imaging (Pearl and Magaritz, 1993). In this
technique, simultaneous imaging of several sequential slices (planes) are collected, and 3D
information is acquired. The signal intensity (S) obtained by imaging is described by Eq. 1
(Budinger and Lauterbur, 1984).
−𝑇𝑅 −𝑇𝐸
𝑆 = 𝑆0 (1 − 𝑒 𝑇1 )𝑒 𝑇2 (1)

where S0 is proportional to the total number of 1H in the resolved volume element (voxel) in an
image.
In MRI, the contrast in the image is obtained through three mechanisms: (1) T1 recovery, (2)
T2 decay, and (3) proton density. For example, the T1 relaxation curve reflects the magnetization
in the longitudinal direction after the RF pulse. If substance 1 recovers faster than substance 2,
substance 1 will have a higher signal and appears brighter on a T1 contrast image, while
substance 2 will appear darker. This is the opposite of the mechanism in a T2 contrast image,
where substance 2 will appear brighter than substance 1. Fig. 3 presents an example of T1
recovery for two substances.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 156
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. T1-weighted contrast between two mediums. Substance 1 is for a medium with
higher signal intensity (e.g. DI water mixed with a contrast agent), whereas substance 2 is
for a medium with lower signal intensity (e.g. DI water only). This contrast is of particular
interest when monitoring the flow of water in porous media.

The proton density (PD) is the number of excitable spins per unit volume. Thus, proton
density determines the maximum signal that can be obtained from a given substance (Deka et al.,
2006). The image contrast in a PD image is independent of T1 or T2 relaxation times, but it
depends on the number of protons in the substance. For example, Pearl and Magaritz (1993)
showed that the MRI signal intensity for a solution of Nickel (Ni) and water increased linearly
with the concentration of Nickel. Similarly, Greiner et al. (1997) showed the same linearity for
copper sulfate (CuSO4), as shown in Fig. 4.

Figure 4. Relative intensity of MRI as a function of the contrast agent (tracer)


concentration.

Tracers (contrast agents). In the context of magnetic resonance imaging, the term tracer or
contrast agent refers to a paramagnetic substance that is used to reduce relaxation time T1 of
water, providing higher signal intensity on a T1-weighted image and producing contrast
enhancement. For example, by dissolving 5 mM of a Ni2+ salt in water, the T1 relaxation time of
the water protons is reduced from about 3 to 0.3 sec at 25 °C. T2 is also reduced to approximately

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 157

the same value (Pearl and Magaritz, 1993). Tracers are specifically used to monitor the flow
process inside water-saturated media (Haber-Pohlmeier et al., 2010). If a tracer-water solution is
introduced to a pure water saturated media, then the injected solution will appear brighter in a T1-
weighted image (Fig. 3), and thus allowing monitoring of the water flow inside the saturated
porous media.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

WATER FLOW IN POROUS MEDIA

MRI has been used for porous media characterization, fluid distribution, solute and fluid
transport, particle and biofilm dynamics, and reactive transport (Werth et al., 2010). The
application of MRI for fluid distribution and solute and fluid transport is of interest in the current
study, so only studies focusing on these two applications are reviewed. Readers are referred to
Werth et al. (2010) for further applications of MRI imaging. Water flow and distribution (Guillot
et al., 1991, Hoffman et al., 1996, Deurer et al., 2002, Haber-Pohlmeier et al., 2021, Amin et al.,
1994), preferential water flow paths (Hoffman et al., 1996, Haber-Pohlmeier et al., 2010), flow
regimes, both convective and molecular diffusion (Pearl and Magaritz, 1993, Greiner et al.,
1997, Oswald et al., 1997), and water content in porous media (Haber-Pohlmeier et al., 2021,
Deurer et al., 2002) have all been studied utilizing MRI. Guillot et al. (1991) obtained the
dispersion coefficients of on Mn2+ tracer in a glass bead medium by fitting 1D profiles with the
solutions of the convection diffusion equation. They compared the flow characteristics for both
gravitationally stable and unstable flow and showed that the 1D profiles for both types of flow do
not differ, but they do in 3D as inferred by the MRI results. Deurer et al. (2002) obtained the
mean velocities and flow rates of water in a glass bead medium and found a good match with
those obtained from a conventional breakthrough experiment. Pearl and Magaritz (1993)
conducted experiments on a sand box filled with two layers of sand; one layer was saturated with
water and the other was saturated with water and NaCl. The two layers were separated by a thin
barrier, which was then removed to study diffusion and convective transport mechanisms. They
showed that the velocity profiles are linearly proportional to the difference in densities between
the two sand layers. They also estimated the permeability from the MRI data and found it to
agree with the measured permeability of the used sand. Greiner et al. (1997) conducted
experiments on flow characteristics of copper sulfates (CuSO4). They compared pore water
velocity, tracer concentration, and cloud propagation from both MRI and outflow breakthrough
curves (OBC), and emphasized on the vantage of MRI in providing 3-D insights on different
flow regimes, rather than an average 1-D flow profile as obtained by OBC.
Moreover, Pohlmeier et al. (2010) studied change in the water content of soils due to root
water uptake. They calculated the water content maps by biexponential fitting of the multi-slice
echo train data and normalization on reference cuvettes filled with glass beads and 1 mM NiCl2
solution. They concluded that the largest water content changes occur in the neighborhood of the
roots in the upper layer of the soil. MRI has been also used for 3D mapping of bacterial density
distribution and tortuosity (Olson et al., 2005, Olson et al., 2004, Sherwood et al., 2003).
Sherwood et al. (2003) attached ferrofluid nanoparticles to bacterial cells and showed that the
bacterial movement can be described by an effective random motility coefficient, consistent with
Brownian diffusion of a nonmotile colloid. Furthermore, MRI has been used to quantify the
velocity of water flow within porous media (Kutsovsky et al., 1996, Manz et al., 1999, Mantle et
al., 2001, Okamoto et al., 2001). Manz et al. (1999) compared the velocity profiles obtained from
MRI to those obtained from Lattice-Boltzmann velocity maps. They showed a quantitative

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 158

agreement between MRI-measured velocity and lattice-Boltzmann simulations for different


Péclet (Pe) and Reynolds (Re) numbers. Mantle et al. (2001) showed that velocity profiles in
alumina pellets are highly heterogeneous, meaning that the majority of the fluid flow occurs
through certain pores. Kutsovsky et al. (1996) applied a flow encoding pulse sequence to
measure the axial velocity profile in a 6 mm glass beads packed in a column. They showed that
water velocities increase toward the center of the pores in a parabolic fashion. Greiner et al.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(1997) utilized a whole-body MR scanner and a 2D spin echo pulse to image copper sulfates
flowing through both homogeneous and heterogeneous glass beads. First and second moment
calculations were used to compute the pore water velocities. The calculated velocities were
consistent with the input values.

PRELIMINARY MRI SCANS

Materials. Badger sand was used for MRI imaging. This sand had less than 0.1% fines as
shown in the particle size distribution curve in Fig. 5. The coefficient of uniformity (Cu) and
curvature (Cc) for this sand were 2.83 and 1.66, respectively, and is thus classified as poorly
graded sand (SP) according to the Unified Soil Classification System (USCS).
Methods. A 77.8 gm of dry badger sand (Fig. 6a) was placed in a polyvinyl chloride (PVC)
container with an outer diameter (OD) of 41.5 mm, height of 53.2 mm, and wall thickness of 3.7
mm. A 12.6 gm of deionized (DI) water were then added to the dry sand (i.e., water content, wc,
of 16.2%). Upon adding water, air voids were expulsed from the specimen by tapping it until no
further air bubbles were visually observed. The specimen was then moved carefully and placed
in the MRI scanner as shown in Fig. 6b and 6c, with an 86 mm volume coil as the transmit coil
and a 20 mm surface coil as signal receiver (Fig. 6b) sited on the top of the container.

Figure 5. Particle size distribution for the badger sand used in this study

MRI Imaging. In this study, the MRI experiments were performed at Farlin Biomedical
Research Institute (FBRI) at Virginia Tech Carilion (VTC), utilizing a multipurpose 9.4 Tesla
high field MR scanner with a high power gradient amplifier, and a spatial resolution as small as
50 m. The gradient strength of the instrument is 660 mT/m, and the maximum linear slew rate
is 4570 Tm/s.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 159
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 6. Experimental setup for the MRI experiment on a saturated badger sand. (a)
badger sand test specimen, (b) the specimen before placing in the MRI scanner, showing
the MRI coil receiver, and (c) showing the setup before starting scanning.

Since water flow monitoring is not the focus of this study, no tracer was used.
Several MRI scanning parameters (image type, TE, TR, bandwidth, slice thickness) were
optimized to increase the imaging quality. Optimization was performed on a coronal section at
the center of the sand specimen. With a field of view of 45x55 mm, both T1-weighted (FLASH)
and T2-weighted (RARE) imaging were attempted, and it was found that for this badger sand, T2-
weighted imaging gives a better quality (less image distortion and better signal-to-noise ratio
(SNR) than T1-weighted imaging but required longer scanning time. Therefore, T1-weighted
imaging was adopted for the badger sand specimen. The echo time and repetition time were
selected as 2.6 ms and 4000 ms, respectively. Moreover, while using a larger slice thickness
reduces the scanning time (a smaller number of images to be collected), a slice thickness of 0.5
mm was selected and used in this study. The 0.5 mm slice thickness corresponds to the mean
particle size (D50) of badger sand. It should be noted that using a smaller slice thickness would
capture the smaller pores within the sand matrix, but this comes with the cost of increasing the
MRI scanning time. On average, the required scanning time for the sand specimen was about 21
minutes 20s for a total of 70 slices

RESULTS AND DISCUSSION

The collected set of images for the badger sand specimen mixed with DI water at 16.2%
water content were processed using MATLAB (MATLAB, 2022), utilizing the volume viewer
and volume segmented modules. The raw data were converted into greyscale binary set of
images, then a three-dimensional volume was reconstructed, as depicted in the color-coded
tomography in Fig. 7, following the image processing technique outlined in Darbari et al. (2017).
The sand pores, or in this case, the pore water, were segmented, and the voxel volumes were then
summed up to obtain the total volume of water inside the sand specimen. The volumetric water
content calculated using the MRI data was 16.0%, which agrees with the initial water content
used to prepare the sand specimen. This indicates that the optimization procedures for the
scanning parameters were successful in capturing all the water signals inside the sand specimen.
It should be noted that the segmentation procedures are subjective and require experience to
select a proper threshold to detect all pore water voxels within the sand matrix.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 160
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 7. 3D color-coded map for the pore water distribution inside the badger sand
specimen

CONCLUSION

This paper reviewed the application of magnetic resonance imaging (MRI) for monitoring of
pore water in porous media. By mapping the movement of the nucleus of interest (in this study,
1
H), magnetic resonance imaging provides information for mapping and monitoring water flow
in porous media in the application of interest in geotechnical and geoenvironmental engineering,
such as seepage, contaminants transport, recharge of aquifers, hydraulic fracturing and internal
and surface erosion. The principles and physics of MRI are presented, and relevant studies
focusing on quantifying water and water flow in porous media are reviewed. Further,
Preliminary MRI scans of badger sand is presented. The water content of a saturated badger sand
sample was calculated both gravimetrically and using a reconstructed 3D tomography for MRI
images of the sand, and both resulted in a similar water content. The authors are currently
investigating different flow regimes in different soil types and the processes of both surface and
internal erosion.

ACKNOWLEDGMENTS

This research was supported by the U.S. Army Corps of Engineers (USACE) under contract
number W912HZ-21-C-0056. The discussions and conclusions presented in this work reflect the
opinions of the authors only.

REFERENCES

Amin, M. H. G., Hall, L. D., Chorley, R. J., Carpenter, T. A., Richards, K. S., and Bache, B. W.
(1994). Magnetic resonance imaging of soil-water phenomena. Magnetic resonance imaging,
12: 319-321.
ASTM. (2016). Standard Test Methods for Measurement of Hydraulic Conductivity of Saturated
Porous Materials Using a Flexible Wall Permeameter. ASTM D5084-16a. West
Conshohocken, PA: ASTM International.
ASTM. (2019). Standard Test Method for Permeability of Granular Soils (Constant Head).
ASTM D2434. West Conshohocken, PA: ASTM International.
Budinger, T. F., and Lauterbur, P. C. (1984). Nuclear magnetic resonance technology for
medical studies. Science 4672: 288-298.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 161

Callaghan, P. T. (1991). Principles of nuclear magnetic resonance microscopy. Oxford


University Press, Oxford, UK.
Darbari, Z., Jaradat, K. A., and Abdelaziz, S. L. (2017). Heating–freezing effects on the pore size
distribution of a kaolinite clay. Environmental Earth Sciences 76: 713.
Deka, K., MacMillan, M. B., Ouriadov, A. V., Mastikhin, I. V., Young, J. J., Glover, P. M.,
Ziegler, G. R., and Balcom, B. J. (2006). Quantitative density profiling with pure phase
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

encoding and a dedicated 1D gradien. Journal of Magnetic Resonance 178: 25-32.


Deurer, M., Vogeler, I., Khrapitchev, A., and Scotter, D. (2002). Imaging of water flow in
porous media by magnetic resonance imaging microscopy. Journal of environmental quality
31: 487-493.
Greiner, A., Schreiber, W., Brix, G., and Kinzelbach, W. (1997). Magnetic resonance imaging of
paramagnetic tracers in porous media: Quantification of flow and transport parameters.
Water resources research 33: 1461-1473.
Guillot, G., Kassab, G., Hulin, J. P., and Rigord, P. (1991). Monitoring of tracer dispersion in
porous media by NMR imaging. Journal of Physics D: Applied Physics 24: 763.
Haber-Pohlmeier, S., Bechtold, M., Stapf, S., and Pohlmeier, A. (2010). Water flow monitored
by tracer transport in natural porous media using magnetic resonance imaging. Vadose Zone
Journal 9: 835-845.
Hoffman, F., Ronen, D., and Pearl, Z. (1996). Evaluation of flow characteristics of a sand
column using magnetic resonance imaging. Journal of Contaminant Hydrology 22: 95-
107.
Kutsovsky, Y. E., Scriven, L. E., Davis, H. T., and Hammer, B. E. (1996). NMR imaging of
velocity profiles and velocity distributions in bead packs. Physics of Fluids 8: 863-871.
Lascialfari, A., and Corti, M. (2007). Basic concepts of magnetic resonance imaging. Springer-
Verlag Italia, Milan.
Manz, B., Gladden, L. F., and Warren, P. B. (1999). Flow and dispersion in porous media:
Lattice‐Boltzmann and NMR studies. AIChE journal 45: 1845-1854.
MATLAB. (2022). version 9.12.0 (R2022a), Natick, Massachusetts: The MathWorks Inc.
Mantle, M. D., Sederman, A. J., and Gladden, L. F. (2001). Single-and two-phase flow in fixed-
bed reactors: MRI flow visualisation and lattice-Boltzmann simulations. Chemical
Engineering Science 56: 523-529.
Nestle, N., Baumann, T., and Niessner, R. (2002). Magnetic Resonance Imaging in
Environmental Science. Environmental Science and Technology 36: 154A-160A.
Okamoto, I., Hirai, S., and Ogawa, K. (2001). MRI velocity measurements of water flow in
porous media containing a stagnant immiscible liquid. Measurement Science and Technology
12: 1465.
Olson, M. S., Ford, R. M., Smith, J. A., and Fernandez, E. J. (2004). Quantification of bacterial
chemotaxis in porous media using magnetic resonance imaging. Environmental science &
technology 38: 3864-3870.
Olson, M. S., Ford, R. M., Smith, J. A., and Fernandez, E. J. (2005). Analysis of column
tortuosity for MnCl2 and bacterial diffusion using magnetic resonance imaging.
Environmental science & technology 39: 149-154.
Oswald, S., Kinzelbach, W., Greiner, A., and Brix, G. (1997). Observation of flow and transport
processes in artificial porous media via magnetic resonance imaging in three dimensions.
Geoderma 80: 417-429.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 162

Pohlmeier, A., Vergeldt, F., Gerkema, E., Van As, H., Van Dusschoten, D., and Vereecken, H.
(2010). MRI in soils: determination of water content changes due to root water uptake by
means of a Multi-Slice-Multi-Echo Sequence (MSME). The Open Magnetic Resonance
Journal 3: 69-74.
Pearl, Z., and Magaritz, M. (1993). Nuclear Magnetic Resonance Imaging of Miscible Fingering
in Porous Media Transport in Porous Media 12: 107-123.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Robbins, B. A., and Griffiths, D. V. (2018). “Internal erosion of embankments: A review and
appraisal”. In Rocky Mountain Geo-Conference, Reston, VA. American Society of Civil
Engineers, 61-75.
Sherwood, J. L., Sung, J. C., Ford, R. M., Fernandez, E. J., Maneval, J. E., and Smith, J. A.
(2003). Analysis of bacterial random motility in a porous medium using magnetic resonance
imaging and immunomagnetic labeling. Environmental science & technology 37: 781-785.
Taina, I. A., Heck, R. J., and Elliot, T. R. (2008). Application of X-ray computed tomography to
soil science: A literature review. Canadian Journal of Soil Science 88: 1-19.
Werth, C. J., Zhang, C., Brusseau, M. L., Oostrom, M., and Baumann, T. (2010). A review of
non-invasive imaging methods and applications in contaminant hydrogeology research.
Journal of contaminant hydrology 113: 1-24.
Zurbuchen, B. R., Zlotnik, V. A., and Butler, J. J., Jr. (2002). Dynamic interpretation of slug tests
in highly permeable aquifers. Water Resources Research 38: 7-1.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 163

Pavement Testing Using Nondestructive MASW Approach

Ramdev R. Gohil1; Jyant Kumar2; and C. R. Parthasarathy, Ph.D.3


1
Ph.D. Student, Geotechnical Engineering, Dept. of Civil Engineering, Indian Institute of
Science, Bangalore. Email: ramdevg@iisc.ac.in
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Professor, Geotechnical Engineering, Dept. of Civil Engineering, Indian Institute of Science,
Bangalore. Email: jkumar@iisc.ac.in
3
Director, Sarathy Geotech and Engineering Services, Bangalore.
Email: partha@sarathygeotech.com

ABSTRACT

In this study, the MASW testing was carried out on asphalt pavement using several
accelerometers and geophones to characterize the stiffness profile of the pavement. Park’s phase-
shift method was applied to the experimental data to obtain the dispersion image in the
frequency-phase velocity domain. Typical pavement structures have high stiffness layers, such as
the asphaltic or concrete layer above the underlying less stiff subgrade. The dispersion image for
such a case shows an increasing trend of phase velocity with increased frequency, which
contrasts with the normally dispersive soil profiles. This obtained dispersion curve can be easily
mistaken as a single continuous curve for inversion analysis, leading to errors in the obtained
profile of the pavement, which includes shear wave velocity (VS ) and thickness of each layer (h).
The predominant modes obtained from the fast delta matrix method for assumed pavement layers
are matched with the effective dispersion curves obtained from experiments to obtain the correct
results for properties of each layer. The automatic inversion using Monte Carlo-based global
search algorithm gives a pavement profile that matches the extracted core sample results.

INTRODUCTION

Multichannel analysis of surface waves (MASW) is one of the nondestructive testing (NDT)
methods in the geotechnical engineering that has attracted the attention of many researchers in
last two decades for determining stiffness profiles of different layers of ground and pavements
(Alzate-Diaz & Popovics, 2009.; Lin & Ashlock, 2015; Nazarian et al., 1983; Nazarian &
Stokoe, 1984 ; Olafsdottir et al., 2020; Park, et al., 2001; Ryden et al., 2001; Ryden et al., 2002).
The NDT methods are preferred choice because of their advantages against the conventional
tests which are destructive in nature. Also, a larger area site can be investigated in a short
duration which will give more reliable results the conventional test method which relies on
sample extracted at only few locations. The basic idea behind the MASW test is to use the
dispersion property of the Rayleigh waves. Rayleigh waves are non-dispersive, meaning they
travel at the same velocity at all frequencies when propagating through homogenous and
isotropic half-space. Only when different layers are present in the media, do the Rayleigh waves
become dispersive, meaning different components of frequencies or wavelengths travel at
different phase velocities. As the Rayleigh wave is a surface wave, it is most dominant on the
shallow top surface within a depth equal to one wavelength (Richart et al., 1970). So, the waves
with a larger wavelength will penetrate deeper into the grounds and waves with a short
wavelength will carry information of only a shallow depth equal to its wavelength. Hence by

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 164

generating the surface wave of a wide range of wavelengths or frequencies, different depths of
layers of soil can be explored.
MASW procedure consists of three main steps: (i) Collection of accurate field data (ii)
Dispersion analysis which will produce a dispersion image (phase velocity vs. frequency plot)
(iii) Inversion analysis to obtain the properties of individual layers by comparing results from
experimental data and numerical results obtained from forward modeling of wave propagation.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

All the steps are crucial and are briefly mentioned here. MASW is an extension and more
sophisticated version of the spectral analysis of surface waves (SASW) test, which employs only
two receivers (Nazarian et al., 1983). By using more receivers in MASW tests, a high-resolution
dispersion image can be obtained which contains information about multiple modes, giving more
accurate results and hence less prone to errors in mode matching where the site is not normally
dispersive. For the case of normally dispersive sites where the stiffness or the shear wave
velocity keeps increasing with the depth, the fundamental mode dominates in the field, which
can be easily identified with decreasing trend of phase velocity with the frequency. But for the
case of an irregular soil profile where either a high-velocity layer is present between two soft
strata or a low-velocity layer is present between two stiffer strata, the dispersion curve is not
necessarily the fundamental mode curve but rather a combination of different modes in the
different frequency range (Tokimatsu et al., 1992). Such profiles are difficult to identify
accurately using two receiver systems. In all these cases, the half-space over which different
strata are lying is always considered the stiffest. But in the case of pavements where the top
layers are stiffer such as surface bitumen or concrete layer, than the underlying base, sub-base
and subgrade layers, the stiffness profile decreases. For such inverse cases, wave propagation
becomes a complex phenomenon to understand and analyze. The objective of this study is to
provide fast and reliable method for accurate inversion analysis of pavements in MASW
approach. The experimental setup, dispersion, and inversion analysis used in this study are
explained in the next section.

METHODOLOGY

Experimental Setup: In the first part, a set of six geophones of 4.5Hz natural frequency are
used to record the vertical component of the signal on the asphalt pavement inside the institute
campus. The temperature during the test was recorded at around 25℃. The vertical impact is
delivered by hitting a handheld hammer directly on the pavement surface, which produces high-
frequency content necessary to investigate the shallow top layers of the pavements (Ryden et. al.,
2001). The geophones are usually fixed on the soil site using the spikes which are difficult to
insert on the hard pavement surface. Hence, they are mounted on heavy aluminum disks which
provides stable base and hence good coupling of receivers with the surface. The picture of the
site setup is shown in Figure 1(a). The receiver spacing is varied as 0.25m and 0.5m. The offset
distance is varied as 0.5m, 1m and 2m for the case of geophones. The data is recorded in the
seismograph and the results are analyzed in MATLAB. The maximum sampling rate available
for the geophones is 2048 samples/second.
The field setup and shot gather obtained using geophones are shown in Figures 1(a) and 2(a),
respectively. In the later part, 14 high-frequency piezoelectric accelerometers are used to record
the vertical component of the signal generated by the hammer. Sticky wax is used to mount the
accelerometers on the pavement surface, which can be easily glued and can be removed easily.
This way of coupling is found to be quick and more reliable method (Lin & Ashlock, 2015). The

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 165

spacing between the receivers is kept as 0.1m, while the offset distance is kept as 0.5m. These
parameters are decided based on the experience of the user with many trials and error to get the
best dispersion results and it agrees with recommended values available in literature (Ryden et
al., 2001, Ryden et. al., 2002, Nazarian et. al., 1983, Lin & Ashlock, 2011). Piezoelectric
accelerometers of Dytran 3055B3 & Kistler 8632C5 models are used in a single experiment,
which have a natural frequency of 35kHz & 9kHz, respectively. Generally, the accelerometers
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

are shown to capture signal with much higher frequency content than their natural frequency (Lin
& Ashlock, 2011) and since the amplitude of all signals is normalized in the dispersion analysis,
both types of accelerometers are used together to improve the results of the analysis. The data is
acquired using the NI data acquisition system, which has the highest sampling rate of 204 kHz
for each of the two modules. The sampling rate of 29000 samples/second is used to record the
signal. The recorded data are obtained in the LABVIEW software on the laptop and further
analysis is done in MATLAB. The assembly and site setup are shown in Figures 1(b) and 2(b),
respectively.

(a) (b) (a) (b)

Figure 1. (a) Test setup using 6 geophones; Figure 2. Shot gather obtained using (a) 6
Receiver spacing 0.25m; (b) Test setup geophones with receiver spacing 0.5m &
using 14 accelerometers; Receiver spacing source offset 1m (b) 14 accelerometers with
0.1m receiver spacing 0.1m & source offset 0.5m

Dispersion Analysis: For obtaining the dispersion image from the recorded data, Park's
phase shift transformation technique is used here, which is a common method for obtaining field
dispersion curves (Park et al., 1998). Figures 3(a) shows the best dispersion image obtained from
the set of experiments conducted using the geophones. The best results are found using the
source offset of 1m and receiver spacing of 0.5m. Similarly, Figure 3(b) show the results
obtained for the case of accelerometer with a source offset of 0.5m and receiver spacing of 0.1m.
The geophones used in this study are not designed to capture the required high-frequency signal.
As a result, it can only give accurate results up to 1000 Hz, as shown in Figure 3(a). The
dispersion curve has a wide band, making it difficult to pick one velocity value for any
frequency. It has been found in the literature that this short and limited range of frequency is not
enough to capture the details of the top layers accurately (Park et al., 2001). Figure 3(b) shows

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 166

the accelerometer data's dispersion image. This time the results obtained are more refined than in
the geophone case. The thin red line shows good resolution of using high-frequency receivers.
The frequency range was limited by the highest sampling rate of the data acquisition system by
using 14 receivers. But it can be easily increased by using a better DAQ system. The next
important part of this step is to extract the dispersion curve data from the image as the maxima of
the amplitude are shown as red lines. Manual picking is preferred to avoid picking the spurious
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

modes. These picked single continuous curves do not represent a single modal curve but are
made of contributions from different modes. After the extraction process, the next step of
inversion is carried out as shown in next section.

(a) (b)

Figure 3. (a) Best dispersion image obtained from experimental data using 6 geophones
with receiver spacing 0.5m & source offset 1m; (b) Best dispersion image obtained from
experimental data using 14 accelerometers with receiver spacing 0.1m & source offset 0.5m

Inversion Analysis:

The inversion analysis involves two parts. The first part is to select a forward modeling
method to obtain the dispersion curves for the frequency and velocity of interest. The second part
involves optimization, where best-fit parameters are found by trial and error so that experimental
dispersion data matches well with the theoretical dispersion results. There are several methods in
the literature to obtain these theoretical dispersion curves. The fast delta matrix method (Buchen
& Ben-Hador, 1996a) is used here to calculate the dispersion curves for the assumed layered
media for the pavement. This method is used here because it has been proved to be
computationally superior to the other methods like the stiffness matrix method (Kausel &
Roësset, 1981), Knopoff method (Schwab & Knopoff, 1972), reflection matrix methods (Kennett
& Kerry, 1979) & other methods based on finite difference approach (Kumar, 2011). To carry
out a quick analysis of pavements on the field, such a method is more suitable, which gives
accurate results in a short time. Hence such a method can also be employed for automatic
inversion analysis, where the profile with the least misfit is to be found out from the pool of all
possible values of layer properties. The theoretical formulation of the method is not shown here
and can be found easily in literature (Buchen & Ben-Hador, 1996b). Instead of using
fundamental or multiple modes separately to compare theoretical and experimental dispersion
results, in this study, the predominant modes obtained from the theory for the assumed pavement

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 167

profile are compared with the apparent dispersion curve obtained from the experimental results.
When the site stiffness is not regularly increasing like in the case of pavements, the fundamental
mode will not be the dominant mode in the frequency range of interest. This is the reason behind
using the predominant mode for inversion analysis rather than the usual fundamental mode.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Dispersion curve maxima extracted from the experimental results

These predominant modes are found by many authors using different methods in the past.
Gucunski & Woods carried out a numerical simulation of the SASW test using an axisymmetric
model of a soil system with circular loading where the receivers' responses at some distance are
represented by vertical displacements (Gucunski & Woods, 1992). The authors derived the
vertical displacements in the frequency-wave number domain using Hankel transform and the
derivations were based on the stiffness matrix approach of Kausel. A similar approach was taken
by (Lin et al., 2022) to simulate the MASW setup and calculate the dispersion image for the
assumed layered media and information about offset distance and receiver spacings. Tokimatsu
et al. computed the superposed modes similar to predominant modes where the vertical
displacements of the first 'n' modes are combined to establish the corresponding dispersion curve
associated with the nth superposed mode. Zomorodian & Hunaidi (2006) computed predominant
modes for the SASW method using the maximum flexibility coefficient. Kumar (2011) and
Naskar & Kumar (2017) calculated predominant modes by comparing the vertical displacement
at the surface obtained from the eigenvectors of displacement using the finite difference
approach and thin layer method, respectively.
In this method, initially the multimode dispersion curves are obtained using the fast delta
matrix method, as mentioned previously. From this data, the predominant modes are found using
the following approach. For each frequency value, there will be corresponding multiple values of
phase velocity for different modes. To find which mode will dominate in the field, the global
stiffness matrix, 𝐊, is assembled using the global stiffness matrix for all the velocity values for a
fixed frequency. By assuming that the load is applied only on the free surface in the vertical
direction, the force vector is assumed as 𝐟 = [0 1 0 0 … 0 ]𝑇 to make other components of force
vector zero. The equation 𝐟 = 𝐊 𝐮 is solved to find the displacement vector 𝐮 by inverting the
stiffness matrix method. The vertical displacement at the top surface is compared for all the
values of velocities for the same frequency. Whichever mode that gives maximum vertical
displacement at the top surface is taken as the predominant mode for that frequency. The process
is repeated for all the frequency values to find predominant modes. These predominant modes

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 168

for different assumed layered media are then compared with the manually extracted apparent or
effective dispersion curve from the experimental results. By doing either the manual inversion or
automatic inversion, the profile that matches fine with the experimental results is considered a
suitable profile.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Misfit = 1.85 %

Best matching curve

Experimental curve Best fit VS profile

Figure 5. Results of automatic inversion analysis for the pavement

Figure 6. Predominant modes obtained from forward modeling in comparison with the
experimental dispersion image

In the past, the method used for inversion analysis was to correlate the depth of investigation
with the maximum and minimum observed wavelength in the SASW analysis of pavements and
soils (Nazarian & Stokoe, 1984). The maximum depth of investigation can be approximated as

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 169

one-third or half of the maximum wavelength observed, whereas the thickness of the top layer is
limited to one-third to half of the shorted wavelength. This method is suitable only when the
fundamental mode is known to dominate in the field, which is possible only when the site is
normally dispersive. In the case of inverse profiles like pavements, where the apparent dispersion
curve is the result of multimodal curves, this approximation is not valid. In this study, a
MATLAB-based open-source software, MASWaves, is used to do the inversion analysis. This
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

tool was developed by Olafsdottir et al., for the inversion analysis based on fundamental mode
(Olafsdottir et al., 2020). Here, this code has been modified to do the inversion based on
predominant modes rather than only the fundamental mode. It uses Monte Carlo-based global
search algorithm to randomly search for the values of shear wave velocity and thickness of the
layers within the defined range. The number of layers in the system is predefined along with the
Poisson's ratio and the density of each layer. These two factors are known to affect the dispersion
results to the least and hence to reduce the number of unknown parameters, these are assumed a
priorly (Nazarian & Stokoe, 1984). The Monte Carlo-based global search method avoids getting
trapped in local minima of the misfit function as in the case of local search methods like
Levenberg-Marquardt-based algorithms, which assumes linearity between the model parameters
and the experimental data. Although this optimization algorithm is highly non-linear and
provides non-unique solutions, it performed satisfactorily in the past for various soil profiles.
The details of the algorithm can be found in the literature (Olafsdottir et al., 2020). Here, only
brief information is given.

RESULTS AND DISCUSSION

A two-layer system resting on half-space is assumed for the analysis. Typically the
pavements are designed as a two layer system of surface course and base course resting on
subgrade of soil. The initial parameters for inversion of pavements are chosen based on typical
pavements profile where the shear wave velocity is expected to decrease with the depth. The
nominal values of shear wave velocity for asphalt layer are assumed for first two layers along
with the thickness values and are given in Table 1. The value of Poisson’s ratio and density of
each layer were assumed close to the expected value and as assumed in the literature (Lin &
Ashlock, 2011). The layer thickness was found out later from the core sample only to verify the
results of the inversion analysis. In practice to avoid the core sample extraction, the layer
thickness as well as shear wave velocity must be found from the inversion analysis. Hence this
both parameters are varied which affects the dispersion curve most significantly. The search-
control parameters for VS and h for all layers are kept as 5%. During the inversion process, these
parameters specify the range of VS and h values, respectively, that the algorithm can sample in
each iteration. The value of search control parameters is suitably selected to get a optimum range
of search while estimating new parameters in the analysis based on the experience (Olafsdottir et
al., 2020). The maximum number of iterations (Nmax ) is kept fixed as 1000. The misfit function,
which represents % error (𝜀) between theoretical and experimental dispersion curve, is defined
as:
𝑄
1 √(𝑉𝑜𝑏𝑠 − 𝑉𝑡ℎ )2
𝜀= ∑ × 100 (%)
𝑄 𝑉𝑜𝑏𝑠
𝑞=1

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 170

where Q is the number of points extracted from the dispersion curves; Vobs & Vth are the
observed and theoretical phase velocity, respectively.

Table 1. Initially chosen parameters for the inversion analysis


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

VS ρ
ν h (cm)
(m/s) (kg/m3)
2000 0.3 1800 10
1800 0.3 1800 20
400 0.35 1600 -

Table 2. Final parameters from the inversion analysis.

VS ρ E
ν h (cm)
(m/s) (kg/m3) (GPa)
1350 0.3 1800 6.1 (5) 8.53
1157 0.3 1800 15.5 (15) 6.26
327 0.35 1600 - 0.462

The results obtained from these inputs are shown in Figure (5). The minimum value of misfit
obtained after the completion of the search is 1.85% which is considered a good approximation
for this case. The final values of phase velocity and thickness of each layer are shown in Table 2.
The value in the bracket indicates the true value of thickness found using extracted core sample.
Figure (6) shows the comparison of predominant modes along with all the modes obtained from
forward modeling for the best fitting parameters. It also shows the experimental dispersion image
obtained from the field along with the extracted effective dispersion curve in open rhombus
shape marker. All different modes are shown with black lines where as the predominant modes
are shown with filled circle in green. It can be noticed that initially, the experimental trend
follows the fundamental dispersion curve up to a frequency of 5000Hz and then jumps to the first
higher mode and matches closely with the first higher mode. Hence using the fundamental mode
alone would give erroneous shear wave velocity profile. The total amount of time taken for
inversion analysis for 1000 iterations is less than 60 seconds on a laptop with 8 GB RAM, which
shows the computational efficiency of this method. The core sample extracted from the field
reveals several layers of asphalt mixed with fine aggregates. It was known that the roads inside
the campus are often overlayed with thin layers of bitumen mix periodically over the years. The
actual core sample was identified with a 5cm top layer having an asphalt-fine aggregate mix in
good condition, followed by two more layers of asphalt mix of 5cm each. The last layer was the
most deteriorated, which has a thickness of around 5cm resting on the sandy subgrade. The total
thickness of the sample extracted was approximately 20-21cm which matches with the total
thickness of top two stiff layer as obtained from the inversion analysis. The elastic modulus can
be determined from the well-known equation: 𝐸 = 2(1 + 𝜈)𝜌𝑉𝑆 2 and the value for the same is
shown in Table 2. The results of the shear wave velocity obtained are found to be around
expected value of the shear velocity of the asphalt layers and soil. From this study only the
thickness of the layers of pavements could be verified and shear wave velocity could not be
verified as no other tests were conducted on the extracted core sample.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 171

CONCLUSIONS

The MASW tests are done on the asphalt pavement using geophones and accelerometers. As
expected, the low-frequency geophones are not suitable for testing pavements as it needs to
record the high-frequency signal to characterize the shallow top layers of the pavements.
Accelerometers are more suitable for testing pavements. The increasing trend of phase velocity
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

with the frequency can be mistaken as the single-mode curve, which will lead to errors during
the inversion to correctly identify the profile of pavement. Since these apparent curves are a
mixture of different modes dominating in different frequencies, predominant modes calculated
from the forward modeling methods are compared with the experimental results. Using the
predominant modes, automatic inversion is carried out using the Monte Carlo-based global
search algorithms which provides satisfactory results in a short time that makes this method more
suitable for pavement testing.

ACKNOWLEDGEMENTS

The authors wish to thank Sarathy Geotech & Engineering Services (SGES) & the
Confederation of Indian Industry (CII) for funding the research in this study.

REFERENCES

Ali Zomorodian, S. M., and Hunaidi, O. (2006). Inversion of SASW dispersion curves based on
maximum flexibility coefficients in the wave number domain. Soil Dynamics and
Earthquake Engineering, 26(8), 735–752.
Alzate-Diaz, S. P., and Popovics, J. S. (2009). Application of MASW to characterize pavement
top layers. Non-destructive testing in civil engineering, NDTCE’09, France.
Buchen, P. W., and Ben-Hador, R. (1996b). Free-mode surface-wave computations. Geophysical
Journal International, 124(3), 869–887.
Gucunski, N., and Woods, R. D. (1992). Numerical simulation of the SASW test. In Soil
Dynamics and Earthquake Engineering (Vol. 11).
Kausel, E., and Roësset, J. M. (1981). Stiffness matrices for layered soils. Bulletin of the
Seismological Society of America, 71(6), 1743–1761.
Kennett, B. L. N., and Kerry, N. J. (1979). Seismic waves in a stratified half space. In Geophys.
J. R. Astr. Soc (Vol. 57). https://academic.oup.com/gji/article/57/3/557/681409.
Kohji Tokimatsu, B., Member, A., Tamura, S., and Kojima, H. (1992). Effects of multiple modes
on Rayleigh wave dispersion characteristics. J. Geotech. Eng., 118(10): 1529-1543.
Kumar, J. (2011). A study on determining the theoretical dispersion curve for Rayleigh wave
propagation. Soil Dynamics and Earthquake Engineering, 31(8), 1196–1202.
Lin, S., and Ashlock, J. C. (2011). A study on issues relating to testing of soils and pavements by
surface wave methods. Review of Progress in Quantitative Nondestructive Evaluation AIP
Conf. Proc. 1430, 1532-1539 (2012).
Lin, S., and Ashlock, J. C. (2015). Comparison of MASW and MSOR for surface wave testing of
pavements. Journal of Environmental and Engineering Geophysics, 20(4), 277–285.
Lin, S., Yi, T. H., Ashlock, J., and Gucunski, N. (2022). Forward modeling of Rayleigh surface
waves for analytical characterization of dominant dispersion trends. Earthquake Engineering
and Structural Dynamics, 51(1), 240–255.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 172

Naskar, T., and Kumar, J. (2017). Predominant modes for Rayleigh wave propagation using the
dynamic stiffness matrix approach. Journal of Geophysics and Engineering, 14(5), 1032–
1041.
Nazarian, S., Stokoe, K. H., and Hudson, W. R. (1983). Use of spectral analysis of surface waves
method for determination of moduli and thicknesses of pavement systems. Transportation
Research Record, 38-45.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Nazarian, S., and Stokoe, K. H. (1984). Nondestructive testing of pavements using surface
waves. Transportation Research Record, 993, 67-79.
Olafsdottir, E. A., Erlingsson, S., and Bessason, B. (2020). Open-source MASW inversion tool
aimed at shear wave velocity profiling for soil site explorations. Geosciences (Switzerland),
10(8), 1–30.
Park, C. B., Ivanov, J., Miller, R. D., and Xia, J. (2001). Seismic investigation of pavements by
masw method - geophone approach. Symposium on the Application of Geophysics to
Engineering and Environmental Problems.
Park, C. B., Miller, R. D., and Xia, J. (1998). Imaging dispersion curves of surface waves on
multi-channel record. In SEG Technical Program Expanded Abstracts (pp. 1377-1380).
Society of Exploration Geophysicists.
Richart, F. E., Hall, J. R., and Woods, R. D. (1970). Vibrations of Soils and Foundations.
Prentice Hall Inc, Englewood Cliff, NJ.
Rydén, N., Ulriksen, P., Park, C. B., Miller, R. D., Xia, J., and Ivanov, J. (2001). High frequency
MASW for non-destructive testing of pavements—Accelerometer approach. Application of
Geophysics to Engineering and Environmental Problems (pp. RBA5-RBA5).
Ryden, N., Ulriksen, P., Park, C., and Miller, R. (2002). Portable Seismic Acquisition System
(PSAS) For Pavement MASW. In 15th EEGS Symposium on the Application of Geophysics
to Engineering and Environmental Problems (pp. cp-191). European Association of
Geoscientists & Engineers.
Schwab, F. A., and Knopoff, L. (1972). Fast Surface Wave and Free Mode Computations (pp.
87–180).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 173

Reducing Mode Assignment Errors in Surface Wave Inversion for Sites with a Very
Shallow Impedance Contrast Using Love Type Surface Waves

Salman Rahimi1 and Clinton M. Wood, P.E., M.ASCE2


1
Geotechnical Engineer, Arup. Inc., Los Angeles, CA. Email: salman.rahimi@arup.com
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Associate Professor, Dept. of Civil Engineering, Univ. of Arkansas, Fayetteville, AR.
Email: cmwood@uark.edu

ABSTRACT

Over the last several decades, active surface wave testing has received increasing attention in
geotechnical engineering and has become a common method for site characterization. Active
surface wave testing can be performed using Rayleigh- and/or Love-type surface waves.
However, Rayleigh-type surface waves have been the most commonly used wave type among
practitioners in the geotechnical community. This paper provides one case study for a site with a
very shallow impedance contrast where surface wave data processing, particularly mode
assignment, becomes very complex. Both Rayleigh and Love-type surface wave testing was
conducted for the case study. The surface wave data were processed using two different
scenarios: (1) fundamental mode Rayleigh waves only and (2) multimodal inversion using
Rayleigh and Love type surface waves. These results were then compared to the boring to
identify the performance of each approach. The results indicate that the first higher mode
dominated the Rayleigh-type surface wave dispersion data at many frequencies, while the Love-
type surface wave dispersion data were associated with the fundamental mode. Overall, the
results indicate that the inclusion of Love-type surface waves and multimodal inversion can
significantly improve the reliability of the inversion process and the final 1D shear wave velocity
profile for sites with a very shallow impedance contrast. Relying solely on Rayleigh-type surface
wave data for such sites would lead to mode misidentification and data misinterpretation.
Therefore, it is recommended to employ both Rayleigh and Love-type surface waves for sites
with a shallow impedance contrast to avoid data misinterpretation and enhance the reliability of
the site characterization using active surface waves testing.

1. INTRODUCTION

Multichannel analysis of surface wave (MASW) is one of the most popular surface wave
techniques used by practitioners and researchers in the Geotechnical community. This method
was initially developed to retrieve a 1D shear wave velocity (Vs) profile of the subsurface (Park
et al. 1999). However, in recent years, MASW has been utilized for a variety of applications such
as 2D and 3D imaging of subsurface conditions (Pilecki et al. 2017; Rahimi et al. 2021),
landslide investigation (Hussain et al. 2020), health monitoring of earthen hydraulic structures
(Cardarelli et al. 2014), liquefaction assessment (Olafsdottir et al. 2015; Rahimi et al. 2020), and
seismic site classification (Martínez-Pagán et al. 2012; Rahimi et al. 2020; Lin et al. 2021).
The MASW method is an array-based surface wave technique that uses the dispersive nature
of either Rayleigh (MASWR) or Love (MASWL) type surface waves propagating through
subsurface layers. The raw shot-gather data are processed to develop the experimental dispersion
curve typically of the fundamental mode of propagation and used in the inversion process to
retrieve the 1D Vs profile.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 174

Practitioners in the Geotechnical community primarily rely on MASWR for active surface
wave testing regardless of the site’s stratigraphy. However, for sites with a very shallow
impedance contrast (e.g. very shallow bedrock), it has been shown that MASW L can provide a
higher quality experimental dispersion data compared to MASWR (Rahimi et al. 2019; Rahimi et
al. 2021). The blind selection of the surface wave type for subsurface characterization can
sometimes lead to a misprediction of the site’s Vs profile. This misprediction is primarily due to
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

mode misidentification, which is one of the main challenges of active surface wave data
processing.
In this study, one example case study is provided for a site with a very shallow impedance
contrast where MASWR dispersion data are largely dominated by the first higher mode while
MASWL dispersion data are related to the fundamental mode of propagation. Relying solely on
the MASWR for such sites leads to a misprediction of subsurface properties (e.g. depth to a sharp
impedance contrast and Vs of the subsurface layers) due to the mode misidentification. The
paper begins with the site locations and geophysical field measurements conducted for this study.
This is followed by MASW data processing to develop the MASWR and MASWL experimental
dispersion curves. The final 1D Vs profiles generated using two different scenarios (i.e.
fundamental mode Rayleigh waves and multimodal inversion using both Rayleigh and Love type
surface waves) are then compared with the boring results to determine the performance of each
scenario. Finally, some recommendations are provided based on the obtained results.

2. SITE LOCATION AND FIELD MEASUREMENT

The site is located in Northwest, AR, as shown in Figure 1. Active surface wave testing was
conducted using Rayleigh and Love-type surface waves. The goal of active surface wave testing
was to identify the depth to the sharp impedance contrast (i.e. bedrock) and develop a Vs profile
to a depth of at least 30 m below the ground surface to be used for seismic site classification. The
MASWR was performed using a linear array of 24, 4.5 Hz vertical geophones with a uniform
spacing between geophones of 3 m. Rayleigh-type surface waves were generated using a
sledgehammer and an IVI T-15000 vibroseis (Figure 2) source at four source locations of 5 m,
10 m, 20 m, and 40 m from the first geophone in the array. The vibroseis source has been shown
(Wood and Cox 2011) to provide improved data quality and allow deeper Vs profiles than
weaker dynamic sources (e.g., sledgehammers or drop weights). Multiple source offsets are used
to ensure high-quality data, allow uncertainty to be estimated, and ensure near-field effects do
not corrupt the data (Cox and Wood 2011). Data was collected at each source location using the
stepped sine method with frequencies between 5-100 Hz.
The MASWL was conducted using a linear array of 24, 4.5 Hz horizontal geophones with a
uniform spacing between geophones of 2 m. Love-type surface waves were generated using
horizontal sledgehammer blows at source offsets similar to MASWR. At each source location,
five sledgehammer blows were stacked in order to increase the signal-to-noise ratio of the
recorded signals.

3. DATA PROCESSING

The raw shot-gather data from the MASWR and MASWL measurements were processed to
construct the experimental Rayleigh and Love dispersion curves (i.e. frequency versus phase
velocity plot). Dispersion curves were generated using the frequency domain beamformer

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 175

(FDBF) transformation technique (Zywicki and Rix, 1999) which typically provides the highest
quality experimental data (Rahimi et al. 2021). The experimental dispersion curves were
generated by automatically picking the maximum spectral peak in the frequency-wavenumber
domain. Experimental dispersion data points corrupted by near-field effects were removed from
the data (Rahimi et al. 2022). The final dispersion curve was generated by combining the
experimental dispersion data from all source offsets. The final dispersion data from all source
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

offsets was divided into 100 frequency bins from 1 to 100 Hz using a log distribution. The mean
and standard deviation was estimated for each data bin resulting in a mean experimental
dispersion curve with an associated standard deviation. The final experimental dispersion curve
was then used as an input into the Geopsy software package Dinver for the inversion process.
The inversion process was performed for two scenarios: (1) using an experimental dispersion
curve from MASWR as the fundamental mode of propagation and (2) using experimental
dispersion curves from both MASWR and MASWL to perform the multimodal inversion. Similar
parameterization was used for both scenarios to make the results comparable.

Figure 1. Site location.

Vertical and
horizontal
geophones

Figure 2. Active surface wave array and University of Arkansas vibroseis truck.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 176

4. RESULTS AND DISCUSSION

In this section, the results of the inversion process for the two scenarios are first presented
and then compared with the boring information to evaluate the performance of each scenario in
predicting depth to the sharp impedance contrast at the site and constructing a reliable Vs profile
for the site.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

4.1 INVERSION RESULTS USING MASWR EXPERIMENTAL DATA AS THE


FUNDAMENTAL MODE

The results presented in this section are only based on the inversion of the experimental
MASWR data as the fundamental mode of propagation. An example MASWR overtone image for
the 10 m source offset generated using the vibroseis as the source is presented in Figure 3.
Presented in Figure 4 are the MASWR experimental dispersion data along with the 1000 best
theoretical fits from the inversion process and the median of the 1000 best fits. From Figure 4, it
is clear that the fit between the experimental dispersion data and the median theoretical
dispersion curve is quite good, except for the lowest frequency experimental dispersion data
points. Therefore, it can be concluded that all the experimental MASWR data are associated with
the fundamental mode of propagation.

Figure 3. Example MASWR overtone image for the 10 m source offset generated using
vibroseis.

The inverted Vs profiles using the experimental MASWR data as the fundamental mode of
propagation are presented in Figure 5. In this figure, the 1000 lowest misfit, median and
minimum misfit Vs profiles are presented along with the Sigma ln (Vs) for the 1000 lowest
misfit Vs profiles. The median of the 1000 lowest misfit profiles is used as the final 1D Vs
profile of the site.
Based on the goodness of the fit between the experimental and theoretical dispersion curves
in Figure 4 and the low value of the calculated misfit (collective squared error between
experimental and theoretical dispersion curves) parameter (less than 0.5), it can be concluded
that all the experimental MASWR data are associated with the fundamental mode and the final

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 177

inverted 1D Vs profile is reliable for the site. However, as it will be shown in the following
sections, the assumption of the experimental MASWR as the fundamental mode and the
reliability of the final Vs are inaccurate when including the experimental MASWL in the
inversion process.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Experimental dispersion curve of MASWR along with the theoretical fits
assuming the data as fundamental Rayleigh mode.

Figure 5. Inverted Vs profile using MASWR dispersion data. Left shows Vs profiles related
to the 1000 lowest misfit, median and minimum misfit profiles. Right shows Sigma ln (Vs)
for the 1000 lowest misfit Vs profiles.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 178

4.2 INVERSION RESULTS USING MASWR and MASWL EXPERIMENTAL DATA AS


THE FIRST HIGHER AND FUNDAMENTAL MODE

The results of the multimodal inversion using both MASWR and MASWL are provided in
this section. Presented in Figure 6 is an example MASWL overtone image generated using the
sledgehammer as the source. Additionally, shown in Figure 7a and b are the experimental
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

dispersion data from the MASWR and MASWL analyses, respectively, along with the 1000 best
theoretical dispersion curves from the multimodal inversion process and the median of the 1000
best fits for both MASWR and MASWL. In this figure, it is apparent that all the MASWL
experimental data (Figure 7b) corresponds quite well with the fundamental Love wave mode.
However, the MASWR experimental data in Figure 7a is a mixture of the fundamental and first
higher Rayleigh wave modes. From Figure 7a, the low and high-frequency experimental data of
the MASWR fits well with the first higher mode, while the intermediate frequency experimental
data from the MASWR fits well with the fundamental mode of propagation. This contrasts with
the results of the inversion process provided in the previous section using only MASWR data.

Figure 6. Example MASWL overtone image for the 10 m source offset generated using
sledgehammer.

Figure 7. Experimental dispersion curves of MASWR and MASWL along with the
theoretical fits using multimodal inversion.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 179

Overall, a comparison of the fit between the experimental and theoretical data of the
MASWR in Figure 4 and Figure 7 indicates that for sites with a very shallow impedance contrast
relying solely on the MASWR data may lead to mode misidentification and significant
uncertainties in the final 1D Vs profile.
The final 1D Vs profile generated using the multimodal inversion of MASWR and MASWL
data is provided in Figure 8. In this figure, the 1000 lowest misfit, median and minimum misfit
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Vs profiles are presented along with the Sigma ln (Vs) for the 1000 lowest misfit Vs profiles.
The median of the 1000 lowest misfit profiles is used as the final 1D Vs profile of the site.

Figure 8. Inverted Vs profile using multimodal inversion of the MASWR and MASWL
dispersion data. Left shows Vs profiles related to the 1000 lowest misfit, median and
minimum misfit profiles. Right shows Sigma ln (Vs) for the 1000 lowest misfit Vs profiles.

4.3 COMPARISON OF THE VS PROFILES WITH THE BORING INFORMATION

The accuracy of the final 1D Vs profile generated using the two scenarios where the MASWR
experimental dispersion data was assumed as the fundamental mode and the multimodal
inversion of the MASWR and MASWL experimental dispersion data are evaluated through the
comparison with the boring information, and the results are presented in Figure 9. In this figure,
the final Vs profile generated from the inversion of the MASWR experimental dispersion data
(R-Fundamental) is shown in red, and the final Vs profile generated from the multimodal
inversion of the MASWR and MASWL(R&L-Multimodal) is shown in black. The boring was
terminated at a depth of around 10 m since the main goal of the boring was to characterize the
soil above bedrock. Based on the boring information, bedrock was encountered approximately
4.6 m below the ground surface. This matches quite well with the depth to bedrock defined from
the R&L-Multimodal Vs profile. However, depth to the bedrock is underestimated using the R-
Fundamental Vs profile because of the mode misidentification, as mentioned in the previous

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 180

section. This indicates the inclusion of Love-type surface wave and multimodal inversion can
prevent mode misidentification for a site with a very shallow impedance contrast. The average
Vs of the top 30 m (Vs30) for R-Fundamental and R&L-Multimodal scenarios is 672 m/s and 660
m/s, respectively. While the difference between the Vs30 estimated from the two scenarios are
almost negligible for this case study, this could change based on the subsurface stratigraphy and
could become more critical for sites with a Vs30 close to the seismic site classification boundary.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 9. Comparison of the final Vs profile generated using the two scenarios of using
MASWR experimental dispersion data as the fundamental mode and the multimodal
inversion of the MASWR and MASWL experimental dispersion data with the boring.

5. CONCLUSION

This study examined the performance of Rayleigh and Love-type surface waves for site
characterization at sites that consist of a very shallow impedance contrast. In this regard, the
Rayleigh- and Love-type surface waves collected at a site in Northwest Arkansas were processed
using two different scenarios. This includes using the experimental dispersion curve from
MASWR as the fundamental mode of propagation and multimodal inversion of the MASWR and
MASWL experimental dispersion data. The results were then compared to a boring at the site to
evaluate the accuracy of each scenario.
It was shown that for sites with a very shallow impedance contrast, the experimental
Rayleigh dispersion data are dominated at some frequencies by higher modes. However, the
experimental Love dispersion data are dominated by the fundamental mode of propagation at all
frequencies. While only one example case study was presented in this study, this behavior was
observed for several sites with the same subsurface stratigraphy.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 181

Overall, it was observed that for sites with a very shallow impedance contrast where the
mode assignment becomes complex, the experimental Rayleigh dispersion data are dominated by
a higher mode at some frequencies. This means that using only Rayleigh-type surface wave
testing, which is the most common method used in practice for site characterization using active
surface wave testing, can lead to mode misidentification and inaccurate Vs profile for the site.
With the inclusion of the Love-type surface waves and multimodal inversion, one can avoid
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

mode misidentification and considerably enhance the reliability of the site characterization and
Vs profile generated using the active surface wave testing for sites with a very shallow
impedance contrast. Therefore, using both Rayleigh and Love type surface waves is
recommended for sites with a very shallow impedance contrast when using active surface testing.

REFERENCES

Cardarelli, E., Cercato, M., and De Donno, G. (2014). Characterization of an earth-filled dam
through the combined use of electrical resistivity tomography, P-and SH-wave seismic
tomography and surface wave data. Journal of Applied Geophysics, 106, 87-95.
Cox, B., and Wood, C. (2011). Surface Wave Benchmarking Exercise: Methodologies, Results
and Uncertainties, GeoRisk 2011: Geotechnical Risk Assessment and Management (C.H.
Juang et al., eds.), ASCE GSP 224, 845-852.
Hussain, Y., Hamza, O., Cárdenas-Soto, M., Borges, W. R., Dou, J., Rebolledo, J. F. R., and
Prado, R. L. (2020). Characterization of Sobradinho landslide in fluvial valley using MASW
and ERT methods. REM-International Engineering Journal, 73(4), 487-497.
Lin, S., Gucunski, N., Shams, S., and Wang, Y. (2021). Seismic Site Classification from Surface
Wave Data to V s, 30 without inversion. Journal of Geotechnical and Geoenvironmental
Engineering, 147(6), 04021029.
Martínez-Pagán, P., Navarro, M., Pérez-Cuevas, J., García-Jerez, A., Alcalá, F. J., Sandoval-
Castaño, S., and Alhama, I. (2012,). Comparative study of SPAC and MASW methods to
obtain the Vs30 for seismic site effect evaluation in Lorca town, SE Spain. In Near Surface
Geoscience 2012–18th European Meeting of Environmental and Engineering Geophysics
(pp. cp-306). European Association of Geoscientists & Engineers.
Olafsdottir, E. A., Bessason, B., and Erlingsson, S. (2015). MASW for assessing liquefaction of
loose sites.
Park, C. B., Miller, R. D., and Xia, J. (1999). Multichannel analysis of surface waves.
Geophysics, 64(3), 800-808.
Pilecki, Z., Isakow, Z., Czarny, R., Pilecka, E., Harba, P., and Barnaś, M. (2017). Capabilities of
seismic and georadar 2D/3D imaging of shallow subsurface of transport route using the
Seismobile system. Journal of Applied Geophysics, 143, 31-41.
Rahimi, S., Moody, T., Wood, C., Kouchaki, B. M., Barry, M., Tran, K., and King, C. (2019).
Mapping subsurface conditions and detecting seepage channels for an embankment dam
using geophysical methods: A case study of the Kinion Lake Dam. Journal of Environmental
and Engineering Geophysics, 24(3), 373-386.
Rahimi, S., Wood, C. M., and Wotherspoon, L. M. (2020). Influence of soil aging on SPT-Vs
correlation and seismic site classification. Engineering Geology, 272, 105653.
Rahimi, S., Wood, C. M., Wotherspoon, L. M., and Green, R. A. (2020). Efficacy of Aging
Correction for Liquefaction Assessment of Case Histories Recorded during the 2010 Darfield
and 2011 Christchurch Earthquakes in New Zealand. Journal of Geotechnical and
Geoenvironmental Engineering, 146(8), 04020059.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 182

Rahimi, S., Wood, C. M., and Bernhardt-Barry, M. (2021). The MHVSR technique as a rapid,
cost-effective, and non-invasive method for landslide investigation: case studies of Sand Gap
and Ozark, AR, USA. Landslides, 1-16.
Rahimi, S., Wood, C. M., and Teague, D. P. (2021). Performance of Different Transformation
Techniques for MASW Data Processing Considering Various Site Conditions, Near-Field
Effects, and Modal Separation. Surveys in Geophysics.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Rahimi, S., Wood, C. M., and Himel, A. K. (2022). Practical guidelines for near-field mitigation
on array-based active surface wave testing. Geophysical Journal International, 229(3), 1531-
1549.
Zywicki, D. J. (1999). Advanced signal processing methods applied to engineering analysis of
seismic surface waves (Doctoral dissertation, Georgia Institute of Technology).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 183

Investigating Freeze-Thawing Behavior of Saline Soil Using Electrical Resistivity


Measurement

Rui Liu1; Cheng Zhu, Ph.D.2; John Schmalzel, Ph.D.3; Benjamin Barrowes, Ph.D.4;
Danney Glaser5; Michele Maxson6; and Wade Lein, Ph.D.7
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Center for Research and Education in Advanced Transportation Engineering Systems, Dept. of
Civil and Environmental Engineering, Rowan Univ., Glassboro, NJ.
Email: liurui82@students.rowan.edu
2
Center for Research and Education in Advanced Transportation Engineering Systems, Dept. of
Civil and Environmental Engineering, Rowan Univ., Glassboro, NJ. Email: zhuc@rowan.edu
3
Dept. of Electrical and Computer Engineering, Rowan Univ., Glassboro, NJ.
Email: schmalzel@rowan.edu
4
Cold Regions Research and Engineering Laboratory, US Army Engineer Research and
Development Center, Hanover, NH. Email: Benjamin.E.Barrowes@erdc.dren.mil
5
Cold Regions Research and Engineering Laboratory, US Army Engineer Research and
Development Center, Hanover, NH. Email: Danney.R.Glaser@erdc.dren.mil
6
Cold Regions Research and Engineering Laboratory, US Army Engineer Research and
Development Center, Hanover, NH. Email: michele.l.maxson.th@dartmouth.edu
7
Cold Regions Research and Engineering Laboratory, US Army Engineer Research and
Development Center, Hanover, NH. Email: Wade.A.Lein@usace.army.mil

ABSTRACT

With the accelerating rate of global warming, seasonally frozen ground has been an emerging
issue in the Northern Hemisphere. The freeze-thaw cycling of frozen soils is known to be the
cause of various engineering failures of infrastructure in cold regions. To better conduct
construction in cold regions, ground investigations on soil profile distribution and properties of
frozen soils are essential. Researchers found that electrical resistivity method outperforms
traditional ground surveying methods in frozen soils for their greater convenience and cost-
effectiveness. However, detailed investigation into the relationship between electrical resistivity
of soil and a variety of soil properties, especially properties related to the freeze-thaw process, is
still needed. In this study, a series of laboratory experiments are conducted to determine the
relationship between soil electrical resistivity and soil geotechnical properties such as initial
water content, bulk density, and pore fluid concentration under freeze-thaw conditions. Manually
prepared soil samples undergo artificial freeze-thaw cycles, and electrical resistivity and
temperature values are recorded simultaneously. Measurement results are summarized for curve
fitting, and a statistical model showing relationship between electrical resistivity and temperature
during freezing and thawing is given. When the temperature was above the freezing point,
resistivity has a linear relationship with temperature on a semi-logarithmic plot; when the
temperature was below the freezing point, resistivity has a parabolic relationship with
temperature on a semi-logarithmic plot. Result patterns during freezing are also different from
those during thawing. Impacts from other soil properties are also incorporated in the model
through optimization methods. Theoretical basis of the relationships is discussed to reveal
changes in the structure of coils at different stages of freeze-thaw cycling. The findings of this
study are expected to enhance understandings of frozen soil structure and geophysical surveying
in cold regions.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 184

INTRODUCTION

Permafrost soils, referring to soils that remain frozen for more than two consecutive years,
are predominantly found in the Arctic and sub-Arctic regions. More than 50% of the uppermost
soil layers are seasonally frozen in the Northern Hemisphere (Zhang et al., 2002). Permafrost is
present in North America, in areas such as Greenland, Northern Canada, and Alaska (Harris,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2018). Moreover, frozen soils commonly exist in plateau areas, such as the Qinghai-Xizang
(Tibetan) Plateau, China (Zhao et al., 2004).
Two permafrost mechanisms, frost heave and thaw settlement have imposed numerous
challenges to the design, construction, and maintenance of infrastructure in Arctic regions. Frost
heave occurs from the subsequent phase change of pore water to ice with decreasing
temperature. The volumetric expansion that occurs when water transforms into ice creates
pressure within the soil that ultimately results in heaving or bulging at the soil surface (Kouretzis
et al., 2015). On the other hand, the transformation of ice back to liquid water with increasing
temperatures or human activity can cause excessive settlement (Hong et al., 2014; Sun et al.,
2018). These hazards associated with permafrost have earned increasing attention in recent years,
as the aging infrastructure in the United States becomes more vulnerable and rising temperatures
due to climate change in the Arctic make the frozen ground more unstable.
There are numerous examples of infrastructure damage caused by frost heave and thaw
settlement found in literature. Engineers had to overcome the effects of frost heave during the
construction of the Qinghai-Tibet railway (Li et al., 2013). Solar fields in North America have
been altered by frost heave, as their foundations have shifted, changing the angle of the solar
panels and making the field less efficient (Daee, 2021). Wyman (2009) describes wooden piles
for transmission lines in Alaska being jacked approximately 1-2 m out of the ground due to soil
heaving. Contrarily, thaw settlement occurred around the Russia-China Crude Oil Pipeline
causing over a meter of soil settlement in some locations (Wang et al., 2018). These examples
highlight the importance of characterizing frozen soil in susceptible regions so mitigation
measures can be enacted to protect existing infrastructure.
A leading method of frozen soil identification is by measuring the difference in liquid water
content. This can be accomplished with electrical resistivity measurements of the soil, which are
then translatable to water content (Tang et al., 2018). Tang et al. (2018) found an exponential
relationship between soil temperature and electrical resistivity for silty clay in the laboratory
using Nuclear Magnetic Resonance. A more popular method, and a method more applicable in
the field, is the Wenner Method. This testing procedure uses four aligned electrodes equally
spaced. The voltage drop is measured across the two inner electrodes while the outer electrodes
provide a current, which is then correlated to the apparent electrical resistivity of the soil
(ASTM, 2018). A variation of this method is described by Kim et al. (2014) where the soil
sample is inserted into a cylindrical column. Similar to the Wenner Method, electrodes are
located at the ends of the column to provide current and two electrodes are located on the side
face to measure the resistivity. This is a preferred method because it is less intrusive since probes
do not penetrate the sample, unlike the traditional method.
The use of electrical resistivity measurements to indicate other physical properties of soil has
proven in numerous studies to be effective. Kormine (1997) found that sands with larger pore
sizes had a higher electrical resistivity than sands with smaller pore sizes. Sreedeep et al. (2005)
observed exponential relationships between electrical resistivity and the degree of saturation.
Zhang et al. (2014) showed a 75% decrease in the original electrical resistivity when the salt

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 185

concentration increased from 2.5% to 10% by weight in clays. Previously, scholars have
investigated the relationship between soil temperature, initial water content, and the electrical
resistivity of soil that underwent a freeze-thaw cycle in the laboratory. We found that the
relationship between soil temperature and electrical resistivity could be described with a
sigmoidal function when temperatures ranged from -20°C to 20°C. Also, a hysteresis behavior
was observed where the resistivity was higher during the thawing phase than in the freezing
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

phase at a given temperature.


This study led the research further into the effects of varying water content on electrical
resistivity of soils. Additionally, we investigated the impact of concentration sodium chloride
(NaCl) in pore fluid on electrical resistivity. We also established empirical relationships between
electrical resistivity and temperature of frozen soil specimen.

MATERIALS AND METHODOLOGY

Materials

To investigate the correlation between soil temperature and electrical resistivity, we carried
out a multitude of experiments using clayey sand. The soil used in this study was obtained from a
local contractor in New Jersey. According to the Federal Highway Administration (FHWA,
2006) specifications, this soil can be classified as frost-susceptible, as it contains 19% (by
weight) of fine particles smaller than 0.02mm. The basic physical properties of the soil are
shown in Table 1.

Table 1. Basic properties of clayey sand samples.

Properties Values
Optimum water content wop (%) 11.4
Maximum dry density ρd (g·cm3) 1.9
Plastic limit (PL) 17.7
Liquid limit (LL) 25.93
Plasticity index (PI) 8.23
Void ratio (e) 0.59
Specific gravity (Gs) 2.63
Dominant minerals Quartz and Gismondine
Soil Classification SC (Clayey Sand)
Frost susceptibility level Medium

It can be classified as clayey sand (SC) based on the United Soil Classification System
(USCS). The frost susceptibility level of the soil was determined to be medium following the US
Army Corps of Engineers Engineering Manual USACE EM 1110-3-138 (1984). Soil samples
were oven-dried and crushed to minimize clumping, and were passed through a No. 10 sieve to
remove stones and other organic matter. Soil samples were prepared by homogeneously mixing
dry soil with distilled water to achieve the desired water content followed by compacting the soil
in layers to reach the target bulk density. If NaCl was to be added to the sample, the desired
amount was weighed and stirred into the distilled water before mixing the water into the dry soil.
Two different water contents were chosen: 11% and 19%. Two different salt amounts were
selected (% by weight): 0% and 4%.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 186

Methodology

To simulate a freeze-thaw cycle in the laboratory, a Test Equity Model 115 environmental
chamber was utilized. The chamber is capable of reaching -70°C, well below the freezing point
of water. The chamber is also programmable so samples could be subjected to an entire freeze-
thaw cycle automatically. For our experiments, there were three temperature stages: a freezing, a
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

hold, and a thawing stage. The freezing phase started at room temperature (20°C) and over
several hours the temperature dropped to the target temperature. Once at the target temperature,
it was held for a few hours to allow temperature equilibrium to be achieved between the soil and
air in the chamber. The thawing phase started at the target temperature and warmed to room
temperature over the same time interval as the freezing phase. Each combination of NaCl and
water content was run through a freeze-thaw cycle with the lowest temperature reaching -25°C.
Custom containers were fabricated for the electrical resistivity tests. The enclosed container
provided a transmission path for the electrical current and minimized the evaporative water loss
during tests. This was accomplished with an acrylic cylinder with an inner diameter of 4.5cm,
stainless steel plate electrodes (current injection ends), stainless steel rod electrodes (voltage
measurements), and an apparatus for closing and sealing the sample materials.
Figure 1 provides a schematic of the overall experimental setup. Each soil column was
instrumented with four electrodes (two current providers and two voltage detectors) and a
thermocouple fastened in the center of the specimen. The soil columns were placed inside the
environmental chamber with all connecting wires passing through an access port to the external
measurement system. Temperature data were sent to a PC through a data logger and Wi-Fi
modem once per minute. Resistance data were acquired by a Miller 400D device and were
transmitted to the PC through Bluetooth. A programmed multi-switch was installed to alternate
resistivity measurements between the soil columns every minute. The multi-switch allowed for
simultaneous testing of multiple samples by removing the need to manually connect the
resistivity meter to each soil column for each measurement.

Figure 1. A schematic of the experimental setup.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 187

Resistivity values were calculated from the following equation:


𝐴
𝜌=𝑅 (1)
𝑎

where ρ is the electrical resistivity (Ω·m), R represents the resistance (Ω), A is the cross-sectional
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

area of the soil column (m2) which is 1.59 × 10-3 m2 in our experiments, and a is the inner
electrode spacing (m), which is 0.125 m in our experiments. The experimental setup was
validated as repeatable electrical resistivity results were obtained (with a set of 3 specimens with
the same properties and temperature conditions, coefficient of variation <10%).

RESULTS AND DISCUSSION

Influence of initial water content and NaCl concentration

Figure 2 plots the electrical resistivity versus temperature of each NaCl concentration for the
freezing and thawing stages, respectively.

(a) (b)

(c) (d)

Figure 2. The electrical resistivity measurements with change in soil temperature for the
freezing phase only. The NaCl concentration alternatives are as follows: (a) 0% (freezing),
(b) 4% (freezing), (c) 0% (thawing), and (d) 4% (thawing) by weight.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 188

While there was no drastic difference in the electrical resistivity between the freezing and
thawing phases, for the majority of samples, the resistivity was higher during the thawing phase.
As shown in Figure 2, an increase in initial water content decreases the overall electrical
resistivity of the soil sample. Similarly, an increase in the initial NaCl quantity decreases the
electrical resistivity of the sample (Figure 3). The effect that NaCl has on the soil’s electrical
resistivity is more substantial than the impact of varying water content. Furthermore, there is a
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

large drop in electrical resistivity between the samples with no NaCl and 4% NaCl,
demonstrating the strong influence the presence of salt has on a soil’s conductivity as it
undergoes a freeze-thaw cycle.

(a) (b)

(c) (d)

Figure 3. The electrical resistivity measurements with change in soil temperature for
freezing and thawing stages. The water content alternatives are as follows: (a) 11%
(freezing), (b) 19% (freezing), (c) 11% (thawing), and (d) 19% (thawing) by weight.

The reason of the difference of electrical resistivity values during the freezing and thawing
phases was unfrozen water saturation. Similar hysteresis phenomenon was observed in other
researchers’ studies as well. Slater and Glaser (2003) observed a hysteresis response between
electrical resistivity and water saturation levels for sandy soil. Interpreting our results using the
concept of saturation, we can conclude that the liquid water saturation, or unfrozen water
saturation for electrical conduction was reduced with the increasing ice content as temperature
decreases under sub-freezing conditions. Tian et al. (2014) described hysteresis effects between

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 189

unfrozen water content and temperature in artificial clay-sand mixtures using the nuclear
magnetic resonance (NMR) technique, and found that the unfrozen water content during the
freezing stage was generally higher than that during the thawing stage between ‒6°C and ‒2°C.
Wu et al. (2017) showed consistently lower electrical resistivity during freezing compared to that
during thawing in natural permafrost, and suggested that the resistivity shift might reflect
changes in the distribution pattern of the unfrozen water in the soil matrix. Additionally, field
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

measurements in the Alpine Region by Supper et al. (2014) showed a similar hysteresis
phenomenon with lowered resistivity during freezing.

Fitting of experimental data

Best-fit equations were developed to describe the relationships found in Figures 2 and 3.
While universal equations were not developed at this time to mathematically correlate the water
content and NaCl concentration to electrical resistivity, equations were generated to describe
each combination from the above plots. Generally, there were two relationships observed in the
above data: a linear and a square root function. The general equation, as a result, is a piecewise
function (Eqn. 2). The fitting parameters: k1, k2, C1, and C2 were found through data analysis for
each scenario.

𝑘1 𝑇 + 𝐶1 , 𝑇 ≥ 𝑇0
log𝜌 = { (2)
𝑘2 √𝑏 − 𝑇 + 𝐶2 , −15℃ < 𝑇 < 𝑇0

where ⍴ is the resistivity (ohm-m), T is the temperature (°C), T0 is the freezing point (°C), and b
is a constant. k1, k2, C1, and C2 are fitting parameters.

Figure 4. Example of fitting on the experimental data. Function Part 1 represents unfrozen
region, where T>T0; Function Part 2 Function Part 1 represents frozen region, where T<T0.

The linear relationship was observed for temperatures above the freezing point of water
(0°C) while, conversely, a square root relationship was found for temperatures below 0°C. Eqn.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 190

2 is limited to -15°C because of the accuracy loss witnessed below that temperature, particularly
with the soil samples that had no NaCl added. The b-value of the fitting equation represents
mathematically the x-intercept of the square root function. In the application of this model, it is
approximated to the data point with the highest temperature classified under the square root
equation. While this should theoretically be equal to T0 (0°C), in the laboratory setting, the b-
value varied slightly. It was typically below 0°C during the freezing stage and above 0°C during
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the thawing stage. The fitting parameters shown in the tables below are unique to each
combination of NaCl and water amount added initially to the soil.
Figure 4 shows an example of using the established functions to fit one set of experimental
data. The original data of electrical resistivity and temperature were inputted, and then applied
with GRG nonlinear algorithm to solve the values of fitting parameters. In Figure 4, the fitted
function shows exhibited correlation with experimental results, with R2 value above 90%.

CONCLUSION

In this study, a series of laboratory experiments were conducted to determine the relationship
between soil electrical resistivity and soil geotechnical properties including initial water content
and NaCl concentration under freeze-thaw conditions. We summarized testing results to establish
a statistical model showing relationship between electrical resistivity and temperature during
freezing and thawing. When the temperature was above the freezing point, electrical resistivity
showed a linear relationship with temperature on a semi-logarithmic plot; when the temperature
was below the freezing point, electrical resistivity showed a parabolic relationship with
temperature on a semi-logarithmic plot.

ACKNOWLEDGEMENT

The experiments described and the resulting data presented herein, unless otherwise noted,
were funded under PE 0602784A, Project T53 “Military Engineering Applied Research,” Task
08 under Contract W913E518C0008, managed by the US Army Engineer Research and
Development Center (ERDC). The lab support comes from undergraduate students of Rowan
University, including Mr. Flynt Tuller, Mr. Ryan Eno, Ms. Lauren Mulvihil, Mr. Thai Ho, Mr.
Jacob King, Mr. Matt Fernandes, and Mr. Michael Barasso. The authors would like to thank Mr.
Jim Kang, Mr. Cheyne Bradley, and Mr. Jake Rahm from Rowan University for improving the
experimental setup. The work described in this paper was conducted at Rowan University's
Center for Research and Education in Advanced Transportation Engineering Systems
(CREATEs), Mullica Hill, NJ. Permission was granted by the Director, Geotechnical and
Structures Laboratory to publish this information.

REFERENCES

ASTM. (2018). Standard Guide for Using the Direct Current Resistivity Method for Subsurface
Site Characterization (Rep. No. D6431-18). West Conshohocken, PA: ASTM International
Standards.
Clayton, L. K., Schaefer, K., Battaglia, M. J., Bourgeau-Chavez, L., Chen, J., Chen, R. H., Chen,
A., Bakian-Dogaheh, K., Grelik, S., and Jafarov, E. (2021, May 11). Active layer thickness
as a function of soil water content. Environmental Research Letters. Retrieved May 22, 2022,
from https://iopscience.iop.org/article/10.1088/1748-9326/abfa4c/pdf.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 191

Daee, B. (2021). Solar in winter: Mitigating risk of environmental damage to solar panels. pv
magazine USA. Retrieved May 2, 2022, from https://pv-magazine-usa.com/2021/12/30/solar-
in-winter-mitigating-risk-of-environmental-damage-to-solar-panels/.
FHWA. (2006). (rep.). Geotechnical Aspects of Pavements: Reference Manual. Washington, DC:
National Highway Institute.
Harris, S. A. (2018). Geocryology : characteristics and use of frozen ground and permafrost
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

landforms, First edition. ed. CRC Press/Balkema is an imprint of the Taylor & Francis
Group, Boca Raton, Florida.
Hong, E., Perkins, R., and Trainor, S. (2014). Thaw Settlement Hazard of Permafrost Related to
Climate Warming in Alaska. ARCTIC 67, 93-103.
Kim, Y., Kim, K., Hong, S., and Cho, W. (2014). The Variation of Physical Properties in Frozen
Soils at Various Freezing Temperatures. Retrieved December 13, 2020, from
https://asmedigitalcollection.asme.org/OMAE/proceedings/OMAE2014/45561/V010T07A00
7/270888.
Komine, H. (1997). Evaluation of chemical grouted soil by electrical resistivity. Proceedings of
the Institution of Civil Engineers-Ground Improvement 1, 101-113.
Kouretzis, G. P., Karamitros, D. K., and Sloan, S. W. (2015). Analysis of buried pipelines
subjected to ground surface settlement and heave. Canadian geotechnical journal 52, 1058-
1071.
Li, P., Yu, Q. H., Yan, F. Z., Huang, Y. S., and Cheng, D. X. (2013). Main Problems and
Solutions on Design and Construction of Qinghai-Tibet DC Transmission Project in
Permafrost Region. Applied Mechanics and Materials, 353-356, 515-523.
Liu, R., Offenbacker, D., Zhu, C., Schmalzel, J., Mehta, Y., Barrowes, B., and Lein, W. (2021).
Electrical Resistivity Hysteresis Response of Clayey-Sands under Sub-freezing Conditions
(No. TRBAM-21-03121).
M.C. Miller Co. (n.d.). (rep.). Soil Resistivity Test Kit Manual. Sebastian, FL: M.C. Miller Co.,
Inc.
Schaefer, K., Zhang, T., Slater, A. G., Lu, L., Etringer, A., and Baker, I. (2009). Improving
simulated soil temperatures and soil freeze/thaw at high-latitude regions in the Simple
Biosphere/Carnegie-Ames-Stanford approach model. Journal of Geophysical Research,
114(F2).
Slater, L. D., and Glaser, D. R. (2003). “Controls on induced polarization in sandy
unconsolidated sediments and application to aquifer characterization.” Geophysics, 68(5),
1547-1558.
Sreedeep, S., Reshma, A. C., and Singh, D. N. (2005). Generalized relationship for determining
soil electrical resistivity from its thermal resistivity. Experimental Thermal and Fluid Science
29, 217-226.
Sun, Z. Z., Ma, W., Zhang, S. J., Mu, Y. H., Yun, H. B., and Wang, H. L. (2018). Characteristics
of thawed interlayer and its effect on embankment settlement along the Qinghai-Tibet
Railway in permafrost regions. Journal of Mountain Science 15, 1090-1100.
Supper, R., Ottowitz, D., Jochum, B., Römer, A., Pfeiler, S., Kauer, S., Keuschnig, M., and Ita,
A. (2014). “Geoelectrical monitoring of frozen ground and permafrost in alpine areas: field
studies and considerations towards an improved measuring technology.” Near Surface
Geophysics, 12(1), 93-115.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 192

Tang, L., Wang, K., Jin, L., Yang, G., Jia, H., and Taoum, A. (2018). A resistivity model for
testing unfrozen water content of frozen soil. Cold Regions Science and Technology, 153, 55-
63.
Tian, H., Wei, C., Wei, H., and Zhou, J. (2014). “Freezing and thawing characteristics of frozen
soils: Bound water content and hysteresis phenomenon.” Cold Regions Science and
Technology, 103, 74-81.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

USACE. (1984). (rep.). Pavement Criteria for Seasonal Frost Conditions. Washington, DC: US
Army Corps of Engineers.
Wang, F., Li, G., Ma, W., Mu, Y., Zhou, Z., and Mao, Y. (2018). Permafrost thawing along the
China-Russia Crude Oil Pipeline and countermeasures: A case study in Jiagedaqi, Northeast
China. Cold Regions Science and Technology, 155, 308-313.
Wu, Y., Nakagawa, S., Kneafsey, T. J., Dafflon, B., and Hubbard, S. (2017). “Electrical and
seismic response of saline permafrost soil during freeze - Thaw transition.” Journal of
Applied Geophysics, 146, 16-26.
Wyman, G. E. (2009). Transmission Line Construction in Sub-Arctic Alaska Case Study:
“Golden Valley Electric Association’s 230kV Northern Intertie”. Electrical Transmission
and Substation Structures 2009.
Zhang, D., Cao, Z., Fan, L., Liu, S., and Liu, W. (2014). Evaluation of the influence of salt
concentration on cement stabilized clay by electrical resistivity measurement method.
Engineering Geology 170, 80-88.
Zhang, T., Barry, R. G., Knowles, K., Ling, F., and Armstrong, R. L. (2002). Distribution of
seasonally and perennially frozen ground in the Northern Hemisphere. AA Balkema
Publishers Zürich, Switzerland, pp. 1289-1294.
Zhao, L., Ping, C.-L., Yang, D., Cheng, G., Ding, Y., and Liu, S., (2004). Changes of climate
and seasonally frozen ground over the past 30 years in Qinghai–Xizang (Tibetan) Plateau,
China. Global and Planetary Change 43, 19-31.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 193

Reliability of Shallow Bedrock Depth Determination from HVSR Measurements in Central


Missouri

Chang Chi1 and Brent L. Rosenblad, P.E., M.ASCE2


1
Dept. of Civil and Environmental Engineering, Univ. of Missouri, Columbia, MO
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Dept. of Civil and Environmental Engineering, Univ. of Missouri, Columbia, MO.
Email: rosenbladb@missouri.edu

ABSTRACT

Determining the depth to bedrock is important for a variety of geotechnical engineering


applications such as foundation design and slope stability investigations. Direct determination of
bedrock depth by drilling often provides only a sparce spatial sampling of depths. Geophysical
methods, such as refraction, resistivity, and surface wave measurements, provide reliable
estimates of bedrock depths but require extensive equipment deployment and expertise to
interpret the data. In recent years, the horizontal to vertical spectral ratio (HVSR) method has
found widespread application in seismology, earthquake engineering, and geotechnical
engineering. HVSR requires only a single three-component sensor, uses ambient noise as the
source, and, importantly, requires little expertise to determine the site frequency. The method is
well suited as an inexpensive and non-intrusive complement to geotechnical site investigation
programs. Specifically, the HVSR method can be used to estimate the depth of bedrock. Past
studies have demonstrated the ability to derive region-specific relationships to estimate the depth
to bedrock over depths to several hundred feet. In this study, we investigate the viability of
HVSR measurements to determine the depth of shallow bedrock (tens of feet) over a more
localized area (about 300 acres) in Central Missouri. Relationships between site frequency and
bedrock depth are developed for these conditions, along with prediction bounds. The influence of
variable bedrock conditions on the results is assessed and discussed.

INTRODUCTION

The HVSR method is a single-station geophysical method used to estimate the resonant
frequency of a site from ambient noise measurements. The measurement is performed by
recording ambient vibrations for several minutes using a three-component, portable seismometer.
The ambient noise records are divided into individual time windows and the Fast Fourier
Transform (FFT) is applied to each window to convert from the time domain to the frequency
domain. The horizontal spectra are merged and divided by the vertical spectrum and an HVSR
plot is created for each time window. The HVSR plots are then combined into an average HVSR
plot, as shown in Figure 1. When a strong impedance contrast is present, a clear peak in the
HVSR plot is observed, as seen in Figure 1. The frequency of the HVSR peak is strongly
correlated with the resonant frequency of the site. There are two common interpretations of the
origin of the HVSR frequency peak: surface wave ellipticity (Lachet & Bard, 1994) and
vertically propagating shear waves (Nakamura, 1989), as illustrated in Figure 2. Both
interpretations require the presence of a strong impedance ratio (greater than about 4) between
the overlying layer and base material for a reliable peak to be observed (SESAME, 2004). The
impedance is the product of the shear wave velocity (𝑉𝑠 ) and mass density (𝜌). Since mass

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 194

density does not vary significantly between soil and rock, changes in impedance are due
primarily to changes in 𝑉𝑠 . For the idealized case of a uniform layer over rigid bedrock, the site
frequency (𝑓𝑟 ) can be directly related to the layer thickness (𝐻) and shear wave velocity (𝑉𝑠 ),
as:

𝑓𝑟 = 𝑉𝑠 /4𝐻 (1)
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

For the more practical case where the 𝑉𝑠 of the soil changes with depth, the resonant
frequency is approximately related to the average shear wave velocity (𝑉𝑠,𝐴𝑉𝐺 ) and layer
thickness (𝐻) as:

𝑓𝑟 ≈ 𝑉𝑠,𝐴𝑉𝐺 /4𝐻 (2)

Figure 1. Example HVSR ratio plot and standard deviation (black continuous and dashed
lines, respectively). Colored plots are HVSR from individual time windows. The peak
frequency and standard deviation of the peak frequency are also indicated.

BACKGROUND

The HVSR method has been used in a variety of applications, including earthquake
microzonation studies (e.g. Gallipoli et al., 2010; Stanko et al. 2019), determination of the
average velocity of deep basins (e.g. Rosenblad and Goetz, 2009), and depth to bedrock
estimation (e.g. Ibs-von Seht and Wohlenberg, 1999; Lane et. al., 2008). To estimate the depth to
bedrock using Eq. 2 information on the 𝑉𝑠 profile at the site must be known to calculate 𝑉𝑠,𝐴𝑉𝐺 .

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 195
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Illustrations of the two interpretations of the origin of the HVSR peak in terms of
(a) body waves and (b) surface waves, along with (c) the transfer function for 1D shear
wave propagation and (d) ellipticity plot for Rayleigh wave propagation

Ibs-von Seht & Wohlenberg (1999) suggested an alternative approach of developing a


region-specific, nonlinear relationship between the fundamental resonant frequency (𝑓𝑟 ) and
sediment thickness (𝐻) of the form:

𝐻 = 𝑎𝑓𝑟 𝑏 (3)

where 𝑎 and 𝑏 are unknown regression coefficients. The general form of Equation 3 can be
derived using a depth-dependent shear wave velocity model, as shown in Ibs-von Seht and
Wohlenberg (1999). The advantage of developing relationships in the form of Eq. 3 is that
explicit knowledge or measurement of 𝑉𝑠 is not required. This approach is predicated on
consistent geology in the region of interest. Ibs-von Seht and Wohlenberg (1999), Parolai et al.
(2002) and Hinzen et al. (2004) have published equations relating sediment thickness to site
frequency based on correlations to borings. In addition, other studies such as Delgado (2000)
showed the HVSR method and the power function relationship approach can be used to
effectively estimate the depth to rock with errors of 15% or less.
Many of these previous studies focused on developing relationships for large regions (tens of
square miles) and deep bedrock depths (tens to thousands of feet). In this study we investigate
the reliability of a HVSR relationship developed for the geology of Central Missouri to estimate
shallow bedrock depths (from about 10 to 55 ft) over the localized region of the campus of the
University of Missouri (about 300 acres).

METHODS

Site Descriptions

The geology of the study region consists of glacial drift deposits and loess from the
Pleistocene age. These deposits typically consist of lean or fat clay that may be mixed with sand,

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 196

silt, and gravel. Beneath these deposits is bedrock from the Mississippian and early
Pennsylvanian age, composed of the Cherokee shale and the Burlington limestone. The shale is
found inconsistently in the study area with variable thickness (0 to 40 ft) while limestone is
common at depths of about 10 to 60 ft in the study area.
A total of 65 HVSR measurements were performed at sites widely distributed around the
University of Missouri campus where detailed borehole information was available.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

HVSR Data Collection, Processing, and Interpretation

HVSR measurements were recorded using a 2-Hz, three-component seismometer (Geospace


Model type HS-1 3C Array), as shown in Figure 3a. The seismometer was placed within a few
feet of each boring location, oriented with the horizontal components in the N-S and E-W
directions, and leveled. All measurements were recorded using a four-channel Data Physics
Quattro dynamic signal analyzer and SignalCalc Ace 2.4 software (Figure 3b). The Quattro is
powered directly from a laptop computer and the software was set to record ambient noise over a
10-minute period. In this study, HVSR plot generation and data interpretation for all ambient
noise records were performed using the Geopsy software v3.2.1 (Wathelet et. al., 2020). The
results presented here were developed by combining the horizontal components using the
squared average approach. Processing parameters (Figure 4) were selected and used in
accordance with the recommended values from the Geopsy User Guidelines (2005). Examples of
smoothed HVSR spectra where the frequency peak is clearly evident are presented in Figure 4.

Figure 3. The (a) three-component seismometer and (b) all test equipment used for HVSR
measurements in this study.

RESULTS

The frequency of the highest peak and the standard deviation of the frequency of the peak
value were identified from each of the 65 HVSR plots. The data were first interpreted by plotting
the HVSR frequency values versus the depth to the first soil/rock interface (regardless of rock
type), as identified by the descriptions in the boring log. As shown in Figure 5, a clear
relationship of decreased frequency with increased depth to rock was observed, as expected. The
power function relationship of Eq. 3 was used to fit a curve to all 65 data points. Residual values

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 197

between the fit and measured points were also calculated and plotted, as shown at the top of
Figure 5. Larger residuals were observed for the low frequency range of 5 to 10 Hz (deeper
bedrock) while smaller residuals were observed at frequencies above 10 Hz (shallower bedrock).
Overall, the average error between the predicted and observed depths was 23%, however, errors
as large as 110% were observed for some of the points. Two separate 90% prediction bounds
were calculated for the low and high frequency regions, as shown in Figure 5. The prediction
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

bounds, at low frequencies were large (±10 ft) indicating that this relationship would be of
limited use for most practical cases.

Figure 4. Processing parameters (left) and two sample HVSR plots showing a narrow peak
(middle) and broader peak (right).

Figure 5. HVSR frequency versus depth to first bedrock encountered for 65 HVSR
measurements.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 198

The data presented in Figure 5 were next sub-divided into two categories based on bedrock
conditions, namely: soil over limestone and soil over shale over limestone. Of the 65 sites, 20
were performed where soil was directly over limestone (i.e. no shale was present). These data
along with the power function fit and 90% prediction bands are shown in Figure 6. This subset of
the data yielded much smaller residual values when compared to the complete data set, with an
average error between predicted and measured depths of 12% and a maximum error of 30%. The
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

90% prediction bounds were again calculated for low and high frequency regions, but in this case
they were nearly the same.

Figure 6. HVSR frequency versus depth to limestone bedrock for 20 HVSR measurements
where no shale was present in the profile.

Data from the remaining 45 profiles that had a soil/shale interface, are presented in Figure 7
with the frequency plotted versus the depth to the shale interface. The power function fit is again
plotted along with residual values and separate 90% prediction bounds for high and low
frequency regions. It is clear from this plot that the large residuals observed in Figure 5 are
associated with site locations where shale was present in the profile. A comparison of the power
function fits for the soil/limestone and soil/shale/limestone cases are shown in Figure 8.

DISCUSSION

The results from the HVSR measurements from all locations show the expected power
function relationship with depth to bedrock, but with a large degree of scatter, particularly at low
frequencies (Figure 5). The errors associated with the bedrock depth prediction are large enough
that the relationship would be of limited practical use in most cases. However, when the data are

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 199

categorized by bedrock conditions, it is apparent that the large residuals are primarily associated
with the presence of shale in the profile. When the profile consists of only soil over limestone
(i.e. no shale) a reliable and useful relationship to predict bedrock depth over the range of this
study (10 to 42 ft) can be developed, with average bedrock depth errors of about 12% (Figure 6).
Practical use of this relationship, however, would be limited to cases where it was known that no
shale was present in the profile at the site.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 7. HVSR frequency versus depth to shale bedrock for 45 HVSR measurements.

Figure 8. Comparison of power function fits of soil over limestone (red solid line) and soil
over shale over limestone using depth to soil/shale interface (black dash line)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 200

When shale is present in the subsurface profile, the scatter in the HVSR-bedrock depth
relationship is much greater, with average bedrock depth errors of 23% and maximum errors of
over 100%. One possible explanation for the greater scatter at soil/shale/limestone profiles is that
the 𝑉𝑠 of the soil layers at the sites with shale are more variable than the 𝑉𝑠 of the soil at the sites
without shale. However, based on the similarity in soil conditions observed in the boring logs at
all sites, this is not a likely explanation for the observed scatter.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

A more plausible explanation is that the rock interface that is indicated by the peak in the
HVSR plots is the deeper shale/limestone interface. The impedance contrast at the soil/shale
interface may not be large enough to produce a dominant peak in the HVSR plot. If this is the
case, the shale layer will have a great influence on the average velocity of the soil/shale layers
above the limestone. To investigate this explanation, the sites were divided into categories based
on the thickness of the shale layer. The depth to limestone was not known at seven of the sites
(the borings did not extend deep enough) so only 58 data points were used in these relationships.
The shale layer varied in thickness from 0 to 29 ft around the study area. Four categories were
created: no shale (control), 0.5-ft to 5-ft thick shale, 5-ft to 10-ft think shale, and shale thickness
greater than 10 ft. The data were then plotted versus the depth to limestone and power functions
were fit to the data in each category. The data along with the four power function fits are plotted
in Figure 9. The results show that as the shale thickness increases, the relationships deviate from
the control group with the fits shifting to higher frequencies for thicker shale layers. This is
consistent with expectations as the thicker higher velocity shale layer will increase the 𝑉𝑠,𝐴𝑉𝐺 and
shift the frequency to a higher value (see Eq. 3) than the control (no shale) group. This
observation suggests that the poor performance of the HVSR results for the soil/shale/limestone
profile sites is due to the large change in 𝑉𝑠,𝐴𝑉𝐺 above the limestone interface caused by the
presence of the shale layer. This change in 𝑉𝑠,𝐴𝑉𝐺 can be observed in the fitting parameters
presented in Table 1, where the “a” values increase with the shale thickness.

Figure 9. HVSR frequency versus depth to limestone bedrock for 58 HVSR measurements.
The red solid line indicates the control group of no shale and three different dash-lines
indicated different groups of shale thickness ranges.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 201

Due to the existence of two unknown variables: the 𝑉𝑠 of shale and the thickness of shale, it is
not possible to develop a useful and reliable relationship for the soil over shale over limestone
condition. Therefore, the HVSR relationships developed from this study cannot be effectively
used to estimate bedrock depths when shale is present.

Table 1. Measured Parameters “𝒂" and “𝒃” for each profile group
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Group Name “𝒂” coefficient of the “𝒃” coefficient of the


Power Function Power Function
Soil over limestone 339 -1.288
(0 ft of shale) (Control)
Soil over shale over limestone 216 -1.081
(Shale thickness: 0.5-5 ft)
Soil over shale over limestone 291 -1.143
(Shale thickness: 5.5-10 ft)
Soil over shale over limestone 542 -1.302
(Shale thickness: beyond 10 ft)

CONCLUSION

The results from this study illustrated the limitations of region specific HVSR relationships
to predict depth to bedrock when bedrock conditions are complex. Relationships between HVSR
frequency and depth to bedrock were developed for a localized region in Central Missouri. The
results obtained using all data showed the expected power function relationship between HVSR
frequency and bedrock depth but with large scatter about the fit. It was concluded that the wide
prediction bounds for this general relationship would limit its use for most practical cases of
bedrock depth estimation. The large scatter was found to be associated with the variable
thickness of the shale layer influencing the average shear wave velocity and the associated site
frequency. When sites without shale were considered separately, a more reliable relationship
with reasonable prediction bounds was developed. Therefore, for sites where it is known that
shale does not exist in the profile, this relationship could be used for a reliable estimation of
bedrock depth.

REFERENCES

Delgado, J., Lopez Casado, C., Giner, J., Estevez, A., Cuenca, A., and Molina, S. (2000).
Microtremors as a geophysical exploration tool: applications and limitations. Pure and
applied geophysics, 157(9), pp.1445-1462.
Gallipoli, M. R., Albarello, D., Mucciarelli, M., and Bianca, M. (2011). Ambient noise
measurements to support emergency seismic microzonation: the Abruzzo 2009 earthquake
experience. Boll. Geof. Teor. Appl, 52, pp.539-559.
Hinzen, K. G., Weber, B., and Scherbaum, F. (2004). On the resolution of H/V measurements to
determine sediment thickness, a case study across a normal fault in the Lower Rhine
Embayment, Germany. Journal of Earthquake Engineering, 8(06), pp.909-926.
Ibs-von Seht, M., and Wohlenberg, J. (1999). “Microtremor measurements used to map thickness
of soft sediments.” Bulletin of the Seismological Society of America, 89(1), 250-259.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 202

Nakamura, Y. (1989). A method for dynamic characteristics estimation of subsurface using


microtremor on the ground surface. Quarterly Report of Railway Technical Research, 30, 25-
33.
Lane, J., White, E., Steele, G., and Cannia, J. (2008). Estimation of Bedrock Depth Using the
Horizontal-to-Vertical (H/V) Ambient-Noise Seismic Method. 490-502.
Lachet, C. D., and Bard, P.-Y. (1994). “Numerical and Theoretical Investigations on the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Possibilities and Limitations of Nakamura’s Technique.” Journal of Physics of the Earth, 42,
377-397.
Parolai, S., Bormann, P., and Milkereit, C. (2002). “New Relationships between Vs, Thickness
of Sediments, and Resonance Frequency Calculated by the H/V Ratio of Seismic Noise for
the Cologne Area (Germany).” Bulletin of the Seismological Society of America, 92(6), 2521-
2527.
Rosenblad, B., and Goetz, R. (2010). Study of the H/V spectral ratio method for determining
average shear wave velocities in the Mississippi embayment. 112. 13-20.
SESAME. (2004). “Guidelines for the Implementation of the H/V Spectral Ratio Technique on
Ambient Vibrations: Measurements, Processing, and Interpretation.” WP12-Deliverable
D23.12, European Commission – Research General Directorate.
Stanko, D., Markušić, S., Gazdek, M., Sanković, V., Slukan, I., and Ivančić, I. (2019).
Assessment of the seismic site amplification in the city of Ivanec (NW Part of Croatia) using
the microtremor HVSR method and equivalent-linear site response analysis. Geosciences,
9(7), p.312.
Wathelet, M., Chatelain, J., Cornou, C., Giulio, G., Guillier, B., Ohrnberger, M., and Savvaidis,
A. (2020). Geopsy: A User‐Friendly Open‐Source Tool Set for Ambient Vibration
Processing. Seismological Research Letters 2020, 91 (3), p 1878–1889.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 203

Geotechnical Site Characterization with 3D Ambient Noise Tomography: Field Data


Applications

Yao Wang1; Khiem T. Tran, Ph.D.2; Brady Cox3; and Joseph P. Vantassel4
1
Ph.D. Student, Dept. of Civil and Coastal Engineering, Univ. of Florida, Gainesville, FL.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Email: wangyao@ufl.edu
2
Associate Professor, Dept. of Civil and Coastal Engineering, Univ. of Florida, Gainesville, FL.
Email: ttk@ufl.edu
3
Professor, Dept. of Civil and Environmental Engineering, Utah State Univ., Logan, UT.
Email: brady.cox@usu.edu
4
Research Associate, Texas Advanced Computing Center, Austin, TX.
Email: jvantassel@utexas.edu

ABSTRACT

A new 3D ambient noise tomography (3D ANT) method is presented for geotechnical site
characterization. It requires recording ambient noise wavefields using a 2D surface array of
geophones, from which cross-correlation functions (CCF) are then extracted and directly
inverted to obtain S-wave velocity (Vs) structure. The method consists of a forward simulation
using 3D P-SV elastic wave equations to compute the synthetic CCF and an adjoint-state
inversion to match synthetic and field CCFs for extraction of Vs. Compared to conventional
passive seismic methods using characteristics of Green’s function (GF), the main advantage of
the presented method is that it does not require the energy balance at both sides of each receiver
pair to retrieve the true GF. Instead, the source power spectrum density is inverted during the
analysis and incorporated into the forward simulation to account for source energy distribution
for accurate extraction of Vs profiles. The presented 3D ANT method was applied to 3 h of noise
recordings from an array of 196 geophones placed on a grid with 5 m spacing at the Garner
Valley Downhole Array (GVDA) site in California. The inverted 3D Vs model is found to be
consistent with previous invasive and non-invasive geotechnical characterization efforts at the
GVDA site.

INTRODUCTION

Active-source seismic methods such as multichannel analysis of surface waves (MASW) and
full-waveform inversion (FWI) have become efficient tools for geotechnical site
characterization. While these methods can provide accurate subsurface profiles, they require
low-frequency energy (<10 Hz) for deep investigation (> 20 m in depth). As large, powerful
active sources capable of generating such low-frequency energy are expensive and generally not
available for use on most projects, many have sought to take advantage of the low-frequency
ambient noise already present in the environment for deep site characterization.
Besides the conventional anbient noise methods based on dispersion characteristics, full-
waveform inversion (FWI) of cross-correlation functions (CCF) of noise fields have been
recently studied. Toward the FWI of CCFs, structural and source kernels were first derived by
Tromp et al. (2010). The practical field applications have been conducted at global and local
scales (De Ridder and Maddison, 2018; Sager et al., 2018). At engineering scales (< 50 m depth),
the 2D ambient noise tomography (2D ANT; Wang et al., 2021) has recently been developed.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 204

This method uses traffic noise fields with known wave propagation direction (along roadway).
The 2D ANT was able to invert CCFs of the traffic noises to extract 2D Vs profiles to detect
roadway sinkholes. Building on our success of the 2D ANT, this study developed a new 3D ANT
method to directly invert CCFs of ambient noise fields for extraction of subsurface 3D Vs
profiles. The main advantage of inverting CCFs is that it does not rely on Green's function
retrieval. Therefore, it doesn't require the energy balance at both sides of each receiver pair to
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

retrieve the true GFs. Instead, the source power spectrum density is inverted during the analysis
and incorporated into the forward simulation to account for source energy distribution for
accurate extraction of Vs profiles. The 3D ANT method's capabilities are evaluated with field
noise data.

METHODOLOGY

The presented 3D ANT method consists of a forward simulation to compute the cross-
correlation function (CCF) and an adjoint-state inversion to match simulated and observed CCFs
for extraction of subsurface structures. 3D P-SV elastic wave equations and their numerical
solutions (Nguyen and Tran, 2018) are used to simulate noise fields and Green's functions
required for computing the CCF during inversion, as discussed in the following sections.

Forward simulation

The CCF 𝐶 𝛼𝛽 between the two signals 𝑠 𝛼 and 𝑠 𝛽 is explicitly given by:

𝐶 𝛼𝛽 (𝑡) = ∫ 𝑠 𝛼 (𝜏)𝑠 𝛽 (𝑡 + 𝜏)𝑑𝜏. (1)

Variables 𝛼 and 𝛽 are indexing the receiver stations. Equation (1) requires performing the
forward simulation for each source location individually to obtain seismograms 𝑠 𝛼 and 𝑠 𝛽 .
However, it is not practical to explicitly simulate the seismograms due to a large number of
sources with unknown locations. Thus, we adopt the implicit approach (Sager et al., 2018, 2020;
Wang et al., 2021) to compute the CCF. The CCF can be formulated via Green's functions as:

1
𝐶 𝛼𝛽 (𝑡) = ∫ ∫ ∫ G(𝒙𝛼 , 𝒙′ , 𝜔) f(𝒙′ , 𝜔) G∗ (𝒙𝛽 , 𝒙′′ , 𝜔) f ∗ (𝒙′′ , 𝜔) exp(𝑖𝜔𝑡) 𝑑Ω′ dΩ′′ 𝑑𝜔. (2)
2𝜋 Ω′′ Ω′

In this equation, 𝒙′ and 𝒙′′ are two arbitrary locations in the 3-D domain Ω. Integrals ∫Ω′ 𝑑Ω′
and ∫Ω′′ 𝑑Ω′′ denote the integration over domain Ω twice, distinctively. G(𝐱 α , 𝐱, t) is the Green's
function with the source located at 𝐱 α , and f(𝐱, t) is the source function. Assuming that the
spatial correlation length of noise sources is shorter than seismic wavelengths, the source terms
can be approximated with a delta function in space and the source power spectrum density (PSD)
𝑆(𝒙, 𝜔) (Wapenaar, 2004; Wapenaar and Fokkema, 2006):

f(𝒙′ , 𝜔) f ∗ (𝒙′′, 𝜔) = 𝑆(𝒙′ , 𝜔)𝛿(𝒙′ − 𝒙′′), (3)

By its definition, the PSD is a field of scalar values that show the spatial location and the
strength of sources. With this approximation, equation 2 becomes:

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 205

1
Cαβ (t) = ∫ ∫ G(𝐱 α , 𝐱, ω) [G∗ (𝐱 β , 𝐱, ω) 𝑆(𝐱, ω)] exp(iωt) d𝐱dω, (4)
2π Ω

and

Cαβ (ω) = ∫Ω G(𝐱 α , 𝐱, ω) [G∗ (𝐱 β , 𝐱, ω) 𝑆(𝐱, ω)]d𝐱. (5)


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Equations (4) and (5) compute the CCF in the time and the frequency domain, respectively.
Using equation (5), the CCF is computed implicitly for a given noise source distribution (all
noise events) instead of individual noise events. We compute the CCF between 𝐱 α and 𝐱 β by
performing the following steps:
1) Run two forward simulations to compute Green's functions G(𝐱 α , 𝐱, ω) and G(𝐱 β , 𝐱, ω)
with sources at 𝐱 α and 𝐱 β .
2) Multiply G(𝐱 α , 𝐱, ω) with the complex conjugate G∗ (𝐱 β , 𝐱, ω) and the noise source PSD
S(𝐱, ω).
3) Sum over all grid points (integration over space 𝐱), and 4) transform the frequency-
domain CCF to the time domain. In this study, the PSD is inverted from measured CCFs,
and S(x,ω) is the same (average value) for all frequencies within a filtering band.

Adjoint-state inversion

The inversion process minimizes the misfit between the observed and the simulated CCFs to
extract the subsurface velocity structures. We define the misfit (residual) between the observed
and synthetic CCFs as:

𝛿𝐶 = 𝐶𝑜𝑏𝑠 − 𝐶𝑠𝑦𝑛 . (6)

The objective function is then defined as the L2-norm of the misfit:

1 1
E = δCT δC = ∑ ∫ dt ∑ δC2 . (7)
2 2
α β

To optimize the objective function, we analyze the three main components that produce the
misfit of CCFs. They are the source signature (source time function), the source power spectrum
density (PSD), the geologic structure (Vs, Vp, and density). The source signature governs the
shape of the CCF waveforms, the PSD defines the location and strength of the noise sources, and
the geologic structure influences the wave propagation. Among these three components, the
source time function estimation is relatively straightforward because it could be implemented in
the data domain without using a geologic model. The source time function is estimated from the
ambient noise records using a deconvolution method (Wang et al., 2021). This method is applied
to all examples for source signal estimation throughout this work. To address the remaining two
components, we invert the PSD and Vs using their sensitive kernels developed by Sager et al.,
2020.
To update the spatial component of the power spectral density distribution, we use the adjoint
techniques to calculate the PSD kernel (Sager et al., 2020):

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 206

𝐾𝑠 (𝑥) = − ∑ ∫ 𝑢† (𝒙)[𝐺 ∗ (𝛼, 𝑥)𝑆(𝑥)]𝑑𝜔. (8)


𝛼=1

The calculation of this kernel consists of two wavefield simulations. 𝑢† is the adjoint
wavefield and 𝐺 ∗ (𝛼, 𝑥)𝑆(𝑥) is the broadcast wavefield of the forward-propagating wavefield
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

recorded at any area where the PSD is positive (where a source exists). The frequency-domain
multiplication is implemented by a time-domain convolution (zero-lag cross-correlation).
Next, the geologic structure refers to an Earth model 𝒎(𝒙), which characterizes the
subsurface material properties (𝑉𝑝 , 𝑉𝑠 , 𝜌). The vector 𝒎(𝒙) controls Green's functions in the
forward simulation equations (4) and (5). To derive the kernel, we use the elastic wavefield
modeling operator as ℒ(∘), where ∘ is a place holder. With this operator, ℒ(𝑢(𝑥)) represents the
forward elastic wavefield simulation in terms of displacement vector 𝒖(𝒙). The time-domain
finite-difference implementation of ℒ(∘) (Nguyen and Tran, 2018) is used for wavefield
simulation throughout this work. The structural kernel with respect to 𝒎(𝒙) is computed as
(Sager et al., 2020):
𝑛

𝐾𝑚 (𝑥) = ∑ ∫[𝑢† (𝒙)ℒ(𝐶(𝒙, 𝛼)) + 𝐶𝛼† (𝒙)ℒ(𝐺(𝒙, 𝛼))]𝑑𝑡 . (9)


𝛼=1

For the first part inside the square bracket, the adjoint wavefield 𝑢† (𝒙) is numerically
computed by injecting the CCF residuals (equation 6) at the receiver locations. At each reference
station 𝛼, the residual 𝛿𝐶𝛼 consists of 𝑛 channels, which are backward propagated
simultaneously to generate wavefield 𝑢† . The operation ℒ(𝐶(𝒙, 𝛼)) denotes the correlation
wavefield propagates from the noise source to the receivers. To calculate this wavefield, we first
inject the source time function at the reference station 𝛼 and record the forward-propagating
wavefield at any area where there is a noise source (with a PSD magnitude >0). We then do
another forward simulation to broadcast the recorded wavefield and multiply the broadcast
wavefield with the PSD to produce the wavefield ℒ(𝐶(𝒙, 𝛼)).
For the second part inside the square bracket, 𝐶𝛼† denotes the adjoint CCF residual wavefield.
It is calculated by:

𝐶𝛼† (𝑥) = ∫ 𝐺𝛼∗ (𝑥, 𝑡)[𝑢†∗ (𝑥)𝑆(𝑥)]𝑑𝑡. (10)

To compute this wavefield, we perform a forward simulation to backward propagate the CCF
residuals 𝛿𝐶 and record the wavefield at any area with a noise source (PSD>0). Then we do
another forward simulation to broadcast the recorded CCF residual wavefield, and multiply the
broadcast wavefield with the PSD to obtain wavefield 𝐶𝛼† (𝑥). The computing of ℒ(𝐺(𝒙, 𝛼)) is
done by injecting the source time function at station 𝛼 and recording the forward-propagating
wavefield. Similar to the first part in the square bracket, the second part is calculated by
performing time-domain convolution of 𝐶𝛼† (𝑥) and ℒ(𝐺(𝒙, 𝛼)).
More specifically, we formulate and compute the gradients for Lamé parameters λ, μ using
strains of the simulated wavefields as:

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 207

𝑇
𝛿𝜆 = ∫ {(𝜀𝑥 + 𝜀𝑦 + 𝜀𝑧 )(𝜀𝑥𝑐+ + 𝜀𝑦𝑐+ + 𝜀𝑧𝑐+ )
0
+ (𝜀𝑥+ + 𝜀𝑦+ + 𝜀𝑧+ )(𝜀𝑥𝑐 + 𝜀𝑦𝑐 + 𝜀𝑧𝑐 )}𝑑𝑡,
𝑇
(11)
𝛿𝜇 = ∫ {(𝜀𝑥 𝜀𝑥𝑐+ + 𝜀𝑦 𝜀𝑦𝑐+ +𝜀𝑧 𝜀𝑧𝑐+ ) + (𝜀𝑥𝑦 + 𝜀𝑦𝑥 )(𝜀𝑥𝑦
𝑐+ 𝑐+
+ 𝜀𝑦𝑥 )
0
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

𝑐+ 𝑐+ 𝑐+ 𝑐+ )
+ (𝜀𝑦𝑧 + 𝜀𝑧𝑦 )(𝜀𝑦𝑧 + 𝜀𝑧𝑦 ) + (𝜀𝑧𝑥 + 𝜀𝑥𝑧 )(𝜀𝑧𝑥 + 𝜀𝑥𝑧
+ (𝜀𝑥𝑐 𝜀𝑥+ + 𝜀𝑦𝑐 𝜀𝑦+ +𝜀𝑧𝑐 𝜀𝑧+ ) + (𝜀𝑥𝑦
𝑐 𝑐
+ 𝜀𝑦𝑥 +
)(𝜀𝑥𝑦 +
+ 𝜀𝑦𝑥 )
𝑐 𝑐 + + 𝑐 𝑐 + +
+ (𝜀𝑦𝑧 + 𝜀𝑧𝑦 )(𝜀𝑦𝑧 + 𝜀𝑧𝑦 ) + (𝜀𝑧𝑥 + 𝜀𝑥𝑧 )(𝜀𝑧𝑥 + 𝜀𝑥𝑧 )} 𝑑𝑡.

1 𝜕𝒖 𝜕𝒖
The strain tensor 𝜺 is computed via the particle displacement 𝒖 as 𝜺𝑖𝑗 = 2 (𝜕𝑥 𝑖 + 𝜕𝑥𝑗 ).
𝑗 𝑖
Notation 𝜀 is the strain of the forward-propagating wavefield ℒ(𝐺(𝒙, 𝛼)), ε+ is the strain of the
adjoint wavefields 𝑢† , εc is the strain of the propagating of correlation wavefields ℒ(𝐶(𝒙, 𝛼)),
and εc+ is the strain of the adjoint correlation wavefield 𝐶 † . Based on the relationships between
variables Vs, λ, μ, and density ρ, the gradient with respect to Vs can be written as:

δVs = −4ρVsδλ + 2ρVsδμ. (12)

Finally, the PSD and Vs are updated iteratively by

𝑆 𝑝+1 = ‖𝑆 𝑝 + 𝜃1 |𝐾𝑠 |𝐿1 ‖,


{ p+1 p (13)
Vs = Vs + θ2 δVs .

The index p denotes the iteration number. The operator |∘|𝐿1 is the L-1 normalization. The
operator ‖∘‖ denotes a PSD magnitude normalization. This normalization neglects the negative
values and maps the non-negative magnitude of the PSD to the range [0,1]. The step length 𝜃1
and 𝜃2 are positive scalers. The gradient 𝛿𝑉𝑠 is normalized by dividing its maximum magnitude.
In this study, we use 𝜃1 equal to 0.05 (5% of the maximum normalized PSD) and 𝜃2 equal to
0.02 (2% of the maximum Vs) of the current model during inversion.

FIELD EXPERIMENT

A field experiment was conducted at the Garner Valley Downhole Array (GVDA) test site
(Figure 1), in California. The site is located in a seismically active region of southern California,
approximately 115 km northeast of San Diego, 150 km southeast of Los Angeles, 7 km east of
the San Jacinto Fault, and 35 km west of the San Andreas Fault (Teague et al., 2018). More
information about this site can be obtained at http://nees.ucsb.edu/facilities/GVDA. Invasive and
non-invasive geotechnical site charachterization efforts performed previously at the GVDA site
include: downhole seismic testing, P-S suspension logging (Stellar 1996), MASW and
microtremor array measurements (MAM) (Teague et al., 2018), and active-source FWI (Fathi et
al., 2016). These investigations indicated that this site comprises 18-25 m of mostly sandy- to
silty-sand alluvium (AL), overlaying decomposed granite (DG) that eventually transitions into
unweathered granite at depths that vary between about 60 – 90 m across the site. The exact
locations for some of these previous seismic tests are displayed in Figure 1 relative to a large 2D
sensor array used to collect data for this study.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 208
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Plan view of the GVDA test site with the locations of several seismic testing
boreholes and sensor arrays indicated. The blue circles represent the 196, 3-component, 5-
Hz nodal staions used for 3D ANT in the present study, which were deployed in a 14 x 14
grid at a uniform spacing of 5 m.

We deployed an array of 196, 3-component, 5-Hz nodal sensors on a 14 × 14 grid with 5-m
uniform spacing on the ground surface at GVDA. We recorded about 28 hours of ambient noises
over two night-time deployments. The array’s center (33.669°N, 116.673°W) is approximately
60 m to the southwest of a rural highway (the Pines to Palms highway). A ten-second example
record of the recorded noise data is displayed in Figure 2a. In this plot, seismograms are shown
in the offset-time style. The offset is defined by the distance to reference station #1. A traffic-
induced surface wave event can be identified from this plot between 3 s and 4 s. This event
traveled through the receiver area and left a trace of linear appearance in the seismogram.
The noise data CCF is computed and inverted following the same steps as those of the
synthetic example. The entire 28-hour recording is filtered through 2-15 Hz bandwidth, and
divided into one-second segments. The CCF between every station pair is calculated for each
segment and sum over all segments. The CCFs for between individual stations and reference
station #1 are shown in Figure 2b. Consistent wavefroms are observed for most of receiver pairs
and clear arrivals can be identified for reference station 1, which it is the closest to the highway
(Figure 1).
For the inversion analysis, the velocity model is set as 160 m × 160 m × 45 m (X × Y × Z)
and discretized into cells of 2.5 m × 2.5 m × 2.5 m. The receiver patch is near the model’s center
with the receiver X-locations from 45 m to 110 m, and the receiver Y-locations from 35 m to 100
m (Figure 3). This setting allows the modeled area covering over 100 m length of the highway
close to the site, benefiting the noise source distribution estimation. Consulting the SASW result
at the site (Fathi et al. 2016), we used a basic 1D initial model with Vs linearly increased from
175 m/s on the ground surface to 500 m/s at the model bottom (50 m depth). The inversion was
run for a total of 45 iterations. The inversion alternated between updating the PSD for 5

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 209

iterations and upadating the models Vs model for 10 iterations. During inversion, Vp was
updated as twice the value of Vs, and 𝜌 was fixed to 1,800 kg/m3.
Figure 3 shows the inverted normalized PSD with a truncated (0~0.5) color scale used for
better showing the traffic noise directions. This result indicates that ambient noise signal arrive
from all directions, but mainly from the highway to the north. The signal power is especially
strong along the shortest path from the road to the receiver array (Figure 2), indicating that
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

traffic-induced surface waves dominate the recorded noise wavefield and the computed PSD is
consistent with the noise sources.

Figure 2. Examples of raw noise data collected at GVDA and experimental CCFs
calculated relative to station #1. (a) A ten-second noise record example. (b) The CCF
displayed in time-sensor-number style, and (c) offset-time style

The inverted 3D Vs model is displayed in Figure 4. It shows softer materials with Vs < ~250
m/s over the top 10-20 m, which agrees quite well with the expected AL-DG interface, with
rapidly increasing Vs at greater depths. The lateral variation of Vs is minimal, although the
thickness of the softest, near-surface material does vary signifcantly across the array. Due to the
distribution of sensors and the PSD, the characterized area is mainly within the sensor array

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 210

(45~110 m in the X-direction and 35~100 m in Y-direction). The initial model was updated very
little outside of the sensor array.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Inverted PSD for the GVDA dataset. The black triangles represent the
geophones.

Figure 4. Inverted 3D Vs model for the GVDA dataset. (a) The final inverted Vs model.
Black triangles represent receivers on the boundary of the sensor patch. (b) XZ plane slice
at Y=85 m. (c) YZ plane slice at X=85 m. (d) XY plane slice at Z=20 m.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 211

To evaluate the 3D ANT algorithm’s performance, we compare waveforms of the observed


and simulated CCFs. The comparison is carried out using station # 61 as a reference, which is
located near the boring log. At this station, the largest station offset is about 65 m (i.e., the
distance to station 196). At the first iteration (Figure 5a), there is an evident gap of arrival-times
between the observed and simulated CCFs. This difference in arrival-times indicates that the
field subsurface has an S-wave velocity slower than the initial model. This time difference is less
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

evident with small station offsets (less than 30 m), implying that the initial model is more
accurate in the shallow depths. At the final iteration, the waveform match is considerably
improved, and the arrival-time difference is small between the observed and simulated CCFs
(Figure 5b). The waveform misfit still exists after the final iteration, mostly due to: (1) the
random noise in the recorded data, (2) the error of source signal estimation, (3) the error of PSD
estimation, and (4) the error of Vs update. Nevertherless, the inverted Vs model produces much
better waveform matches than the initial model, and thus better represents the actual site
conditions.

Figure 5. CCF waveform comparisons for the GVDA dataset relative to station 61 after: (a)
the first inversion iteration, and (b) the last inversion iteration.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 212

As noted above, Stellar (1996) performed P-S suspension logging (PS logging) at the GVDA
site near station # 47 (X=65 m, Y=85 m; refer to Figure 1). To compare the proposed 3D ANT
method to ground truth, we compare the initial Vs model, the inverted Vs model, and the
simplified PS log data at the borehole's location (Figure 6a). The inverted Vs has good alignment
with the PS logging data above 15 m, and correctly reflects the increasing trend of Vs with
depth. The AL-DG interface is estimated around 20 m in the 3D ANT result, which is slightly
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

deeper than that from the borehole PS logging data. This is likely due to the smoothness of the
3D ANT result and the resulting lack of reflected waves at the layer interface. Nevertherless, it
shows the correct trend of velocity increment with depth.

a) b)

Figure 6. Compariosn of 1D Vs profiles from the present 3D ANT inversion and those
obtained previously at the GVDA site: (a) 3D ANT Vs profile from station # 47 at X=65 m,
Y=85 m in comparison to the PS logging Vs profile in a nearby borehole. (b) 3D ANT Vs
profile from station # 154 at X=110 m, Y=50 m in comparison to nearby SASW and active
FWI Vs profiles.

Fathi et al. (2016) performed SASW and active-source FWI at this site. Figure 6b shows the
comparison of 1D Vs profiles from the 3D ANT at a point near station #154 (X=110 m, Y=50 m;
refer to Figure 1) to those of SASW (line 1) and active FWI (at X = 0 m in Figure 15 of Fathi et
al, 2016). The active FWI study provides accurate Vs profiles that agree with the SASW at
depths above 10 m. However, the active FWI does not characterize a stiffer layer below 20-m
depth, mostly likely due to the lack of low frequency signals (2-5 Hz) used in the analysis. In
contrast, the ANT inverted profile is more consistent with the SASW and PS logging results. The
ANT and SASW produce similar results because both method utilize the low-frequency
components (2-10 Hz) to image the deeper structure.

CONCLUSIONS

This paper presents a new 3D ambient noise tomography (3D ANT) method, which analyzes
noise cross-correlation functions (CCF) for characterization of 3D Vs subsuface structure. The
novelty of the presented method is that it does not rely on Green's function retrieval, or require
the noise wavefield to be far-field. Instead, by accounting for noise source distribution directly, it

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 213

inverts the full-waveform CCF for Vs structures. The method’s capability is tested on a massive
traffic-induced noise dataset collected at Garner Valley Downhole Array (GVDA) test site in
southern California. The inverted noise source distribution is consistent with the known ambient
sources (a nearby highway) and the inverted Vs model is consistent with prior invasive and non-
invasive geotechnical site characterization data. Based on the synthetic and field experimental
results, the ANT method is a useful geophysical tool for characterization of 3D soil and rock
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

profile.

ACKNOWLEDGMENTS

This study was financially supported by the National Science Foundation: grants CMMI-
1930697 and CMMI-1931162. The supports are greatly appreciated.

REFERENCES

De Ridder, S. A. L., and J. R. Maddison. 2018. Full wavefield inversion of ambient seismic
noise: Geophysical Journal International, 215, 1215–1230.
Fathi, A., B. Poursartip, K. H. Stokoe II, and L. F. Kallivokas. 2016. Three-dimensional P-and S-
wave velocity profiling of geotechnical sites using full-waveform inversion driven by field
data: Soil Dynamics and Earthquake Engineering, 87, 63–81.
Nguyen, T. D., and K. T. Tran. 2018. Site characterization with 3D elastic full-waveform
tomography: Geophysics, 83, R389–R400.
Sager, K., L. Ermert, C. Boehm, and A. Fichtner. 2018. Towards full waveform ambient noise
inversion: Geophysical Journal International, 212, 566–590.
Sager, K., C. Boehm, L. Ermert, L. Krischer, and A. Fichtner. 2020. Global-scale full-waveform
ambient noise inversion: Journal of Geophysical Research: Solid Earth, 125,
e2019JB018644.
Steller, R. “New borehole geophysical results at GVDA,” UCSB Internal report; 1996.
Teague, D. P., B. R. Cox, and E. M. Rathje. 2018. Measured vs. predicted site response at the
Garner Valley Downhole Array considering shear wave velocity uncertainty from borehole
and surface wave methods: Soil Dynamics and Earthquake Engineering, 113, 339–355.
Tromp, J., Y. Luo, S. Hanasoge, and D. Peter. 2010. Noise cross-correlation sensitivity kernels:
Geophysical Journal International, 183, 791–819.
Wang, Y., K. T. Tran, and D. Horhota. 2021. Road sinkhole detection with 2D ambient noise
tomography: Geophysics, 86, KS123–KS135.
Wapenaar, K. 2004. Retrieving the elastodynamic Green’s function of an arbitrary
inhomogeneous medium by cross correlation: Physical Review Letters, 93, 254301.
Wapenaar, K., and J. Fokkema. 2006. Green’s function representations for seismic
interferometry: Geophysics, 71, SI33–SI46.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 214

Numerical Investigation of Full Waveform Tomography to Identify Anomalous Conditions


and Untreated Zones in Jet Grout Columns

Pourya Alidoust1; Joseph T. Coe, Ph.D.2; Yang Yang, Ph.D.3; Alireza Kordjazi, Ph.D.4;
and Siavash Mahvelati, Ph.D.5
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Dept. of Civil and Environmental Engineering, Temple Univ., Philadelphia, PA.
Email: pouryaalidoust@temple.edu
2
Dept. of Civil and Environmental Engineering, Temple Univ., Philadelphia, PA.
Email: joseph.coe@temple.edu
3
Dept. of Civil and Environmental Engineering, Temple Univ., Philadelphia, PA.
Email: yang.yang0007@temple.edu
4
RIZZO International, Inc., Pittsburgh, PA. Email: alireza.kordjazi@rizzointl.com
5
Vibra-Tech Engineers, Inc., Hazleton, PA. Email: siavashm@vibratechinc.com

ABSTRACT

Development of infrastructure in areas with poor subsurface soil conditions is often


encountered in civil engineering. In these situations, the engineering characteristics of
problematic soils can be improved by ground improvement techniques tailored to site-specific
constraints. Jet grouting is among the most widely used methods to stabilize poor subsurface
soils, particularly for underpinning of existing structures and cutoff walls for excavation support.
During jet grouting, a stabilizing fluid (cementitious or chemical grout) is injected into the
surrounding soil through small nozzles under very high pressure and velocity. The rotary motion
of the injection nozzle creates columnar elements of soilcrete as the nozzle is withdrawn from
the subsurface. The columns are often overlapped to create a wall or prevent horizontal
movement of unstable soils and/or groundwater. However, local variations in the subsurface
conditions can lead to variability in the column diameter. This in turn can cause untreated zones
where the columns cease to overlap. Typical quality assurance efforts rely on real-time recording
of jet grout parameters such as fluid pressure, flow rate, rotational speed, and rate of
withdrawal/insertion. This is combined with a comprehensive test program prior to production
efforts that establishes the jet grout column spacing, overlap, and overall depth geometry,
typically using destructive methods such as excavations and corings of the soilcrete columns.
However, these efforts do not allow for a comprehensive site-wide assessment of the as-
constructed column geometry given the significant spatial variability present in the subsurface of
many sites. Consequently, there exists demand for a high-resolution non-destructive testing
(NDT) method that can reliably identify untreated zones caused by lack of column overlap after
jet grouting. This paper numerically examines the use of full waveform inversion (FWI) of
seismic data as a potential NDT approach for quality assurance of jet grouting efforts. FWI seeks
to optimize the misfit between synthetic waveforms and observed seismic data. An initial model
is subsequently updated by a linearized inversion procedure for optimizing the wave propagation
properties of the subsurface soils. The current study adopts the spectral element method as a
robust forward modeling technique to solve the wave propagation equation and updates the
models using a quasi-newton approximation of the inverse Hessian to lower the computational
cost of the process. The results of the current study highlighted the potential of the FWI
technique to reliably image anomalous conditions in the overlapping jet grout column and detect
any untreated zones.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 215

INTRODUCTION

Suitable sites for construction purposes is a major recurrent challenge influencing urban
development. Often, the underlying soil conditions cannot reliably withstand design loads for
large infrastructure that is commonplace in modern civil engineering practice. In such
circumstances, ground improvement techniques can be used to improve the engineering
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

properties of the problematic soil, assuring adequate performance for design requirements. A
wide range of ground improvement techniques have been developed to modify different
properties of the natural soil (e.g., compressibility, bearing capacity, permeability, etc.) as per the
site-specific needs of different projects. Due to its high versatility, jet grouting has been
implemented frequently in geotechnical engineering, particularly for the underpinning of existing
structures and cutoff walls for excavation support (Croce and Modoni 2007; Burke and Yoshida
2013).
During jet grouting, a stabilizing fluid (cementitious or chemical grout) is injected into the
surrounding soil through small nozzles under very high pressure (up to 60 MPa) and velocity (up
to 300 m/s) to erode, replace and mix with the in-situ soil (Figure 1). The rotary motion of the
injection nozzle creates columnar elements of soilcrete as the nozzle is withdrawn from the
subsurface. The columns are often overlapped to create a wall or prevent horizontal movement of
unstable soils and/or groundwater (Figure 1). However, the final geometry of the jet grout
columns is highly affected by two major sets of factors: (1) construction parameters such as
injection pressure, suspension density, drilling tolerance, and nozzle conditions (Burke 2012;
Nicholson 2014; Galindo Guerreros et al. 2015); and (2) local variations in subsurface
stratigraphy. Issues with jet grout construction can cause untreated zones where the columns
cease to overlap and design resistances of the improved soils are not obtained (Figure 1). Thus,
quality assurance and verification of the design geometry is considered a critical step to
guarantee the successful performance of jet grout columns.
Due to its challenging nature, several techniques have been proposed to validate the diameter
and shape of jet grout columns. Typical quality control efforts rely either on real-time computer
monitoring of jet grout parameters such as fluid pressure, flow rate, rotational speed, and rate of
withdrawal/insertion (Lee et al. 2005; Gazzarrini 2021) or destructive field-test programs such as
excavations and corings of the soilcrete columns (Stark et al. 2009; Bonetto et al. 2018).
However, these efforts do not allow for a comprehensive site-wide assessment of the as-
constructed column geometry given the significant spatial variability present in the subsurface of
many sites. To address the problem, surface or borehole-based nondestructive testing (NDT)
methods have been evaluated as non-invasive, cost-effective, and less time-consuming
approaches (e.g., Cheng et al. 2017; Bonetto et al. 2018; Lin et al. 2020). Although the current
standard-of-practice NDT methods provide beneficial information regarding subsurface
conditions, many suffer from either low resolution, limited coverage, and/or insensitivity to
spatial variations in the subsurface when applied to quality control of jet grout columns.
Consequently, there exists demand for a high-resolution NDT method that can reliably identify
untreated zones caused by lack of column overlap after jet grouting.
The continued improvement in computational resources has promoted advancements in NDT
methods that have ultimately led to better characterization of subsurface conditions. Among
these techniques, Full Waveform Inversion (FWI) has emerged as a robust seismic tomographic
approach to obtain a high-resolution multi-parameter map of subsurface media. This method
offers the potential to overcome the drawbacks of conventional NDT techniques in terms of

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 216

resolution and accuracy. FWI was first proposed by Tarantola (1984) and Mora (1987) and
works to characterize subsurface conditions by numerically simulating wave propagation in a
domain and iteratively minimizing the difference between the synthetic data and recorded
waveforms from the field. In this manner, FWI typically proceeds as a gradient-based local
optimization whereby the minimum of a misfit function is sought. Once the misfit is minimized,
the subsurface model used to numerically simulate wave propagation is assumed to represent
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

actual subsurface conditions in the field. Promising results from FWI have been reported for
NDT purposes in recent studies, including in applications related to characterizing unknown
foundations (Nguyen et al. 2016; Mahvelati and Coe 2019; Kordjazi et al. 2022), evaluating
anomalous conditions in drilled shafts (Kordjazi et al. 2021), and detection of delamination and
rebar debonding in concrete structures (Chen et al. 2022).

(b)

(a) (c)

Figure 1. Ground improvement with jet grouting: (a) schematic of typical construction
process (Bearce et al. 2016); (b) exhumed columns showing example geometries (Mooney
and Bearce 2018; Njock et al. 2018); and (c) untreated zones due to variation in column
geometry (adapted from Mooney and Bearce 2018).

Despite evidence of its high resolution, seismic FWI has not yet been documented to evaluate
anomalous conditions in jet grout columns. The current study aims to numerically explore the
capabilities of FWI when applied in this manner. The spectral element method (SEM) was used
as a robust forward modeling technique to solve the wave propagation equation and model
updates were performed using a quasi-newton method to approximate the inverse Hessian and
lower the computational cost of the process. The results highlighted the potential of the FWI

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 217

technique to reliably image stratigraphy-related anomalous conditions in overlapping jet grout


columns caused by untreated zones.

SUBSURFACE MODEL AND NUMERICAL SIMULATIONS

A two-dimensional subsurface model was developed in this study to simulate a reduction in


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the diameter of overlapping jet grout columns (Figure 2). It was assumed that the model domain
represented a pilot test program over a limited region of the site to evaluate the viability of jet
grouting and the effects of construction parameters (e.g., injection pressure, etc.) prior to
production efforts. Each jet grout column had a diameter of 2.0 m and was embedded 10.0 m into
a two layer soil profile. Near the mid-depth of the subsurface profile is a 3.0 m thick stiffer soil
layer that reduces each column diameter to 1.0 m and causes an untreated soil zone. Wave
propagation properties for each of the materials in Figure 2 are listed in Table 1. The entire
domain was 40.0 m long and 10.0 m deep and was spatially discretized with an average element
size of approximately 0.83 m in both directions. The resulting model contained 576 elements.
Numerical simulations of seismic wave propagation within the model were performed using
the SalvusCompute module of the Salvus software suite (Afanasiev et al. 2019). SalvusCompute
uses a spectral-element method (SEM) (Tromp et al. 2008) formulation for the elastic wave
propagation equations. Boundary conditions for the model consisted of a free surface at the top
and Clayton and Engquist (1977) absorbing boundaries at the bottom/sides of the domain.
Observed waveforms were measured using 20 surface receivers located on both sides of the
overlapping jet grout columns (Figure 3). A total of six source impulse signals were input into
the domain using a 30 Hz vertically-polarized Ricker wavelet. Similar to the receivers, these
source impulse waves were located on either side of the jet grout columns. This layout for
sources and receivers mimicked a typical data acquisition scheme as would be deployed for
surface-based seismic geophysical testing with either surface waves (e.g., multichannel analysis
of surface waves, MASW) or body waves (e.g., seismic reflection/refraction). FWI attempts to
match the entirety of the wavefield record, which means that both surface and body waves are
automatically modeled during analysis. Therefore, FWI could be performed using a combination
of surface/body waves from surface receivers and body waves from downhole receivers. However,
downhole receivers were purposefully avoided in this study to explore the capabilities of FWI
where no drilling/probing alongside the columns is necessary for seismic data acquisition.

Figure 2. Jet grout column true model.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 218

Table 1. Wave propagation properties of the materials in the jet grout column true model.

Primary Wave Shear Wave


Material Density, ρ (kg/m3)
Velocity, VP (m/s) Velocity, VS (m/s)
Soft Soil 1900 200 100
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Stiff Soil + Column


2000 400 200
Anomalies
Jet Grout Column (Intact) 2100 1000 500

DATA PROCESSING

The observed waveforms from wave propagation simulations on the true model were
subsequently used for FWI. The SalvusOpt module was used to control the inversion workflow.
Since the model domain represented a grouting test program, it was assumed that the subsurface
soil conditions would already be well characterized from previous site investigations. In this
context, FWI is intended as a diagnostic tool to evaluate changes in grout column geometry only.
Consequently, the initial model for the inversion workflow could be well constrained with a-
priori knowledge of the stratigraphy and wave propagation properties of the soils (Figure 3).
Uncertainty in the grout column properties was explored by performing three sets of FWI
analyses with different starting models. Each starting model replicated the geometry in Figure 3,
but the initial values for the VP, VS, and ρ parameters of the jet grout column were varied as
follows: (1) No difference from the true model; (2) 10% decrease relative to the true model; and
(3) 20% decrease relative to the true model. These initial models are hereafter referred to as the
0% initial model, 10% initial model, and 20% initial model, respectively. It was also assumed
that the top of the jet grout column would be visible after construction. The diameter at the
surface could be subsequently used to further constrain the inversion workflow by establishing a
region of interest that limited the geometry over which computed gradients were applied to the
model properties (Figure 3).

Figure 3. Initial starting model for FWI with location of sources, receivers, and region of
interest where computed gradients were applied to model properties.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 219

During the inversion, the misfit between the observed waveforms recorded at the receivers
and those from forward modeling was computed using a least-squares norm (L2) function. The
computed gradients were used to update VP, VS, and ρ in the aforementioned region of interest.
The inversion workflow also incorporated a multi-scale approach (Bunks et al. 1995). Initial
iterations of the inversion are performed using a low frequency source wavelet. The resulting
subsurface model is used as the initial model for subsequent iterations with a higher frequency
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

source. This process is repeated multiple times with source wavelets of increasingly higher
frequency. This multi-scale approach has been shown to encourage successive model parameter
estimates to converge towards the global minimum of the FWI optimization problem, thereby
reducing the likelihood of convergence towards local minima caused by cycle skipping (Yao et
al. 2019). In this study, the central frequency of the Ricker wavelet source was 10 Hz, followed
by 15 Hz and 30 Hz, respectively. In total, the inversion workflow terminated after
approximately 30 - 35 iterations depending on the initial model. At the end of the inversion, the
relative misfit was less than 2% of the initial misfit in all cases. In terms of computational
resources, the calculations were submitted to eight cores of the Compute servers of the Temple
High Performance Computing cluster. Compute is an interactive-use server that provides 88
CPU cores [Intel® Xeon Gold 6238 (Cascade Lake) processors] with up to 1.5 TB of RAM and
0.5 PB shared memory. Each iteration of the inversion required approximately six minutes to
complete, resulting in a total inversion run time of approximately three hours.

RESULTS AND DISCUSSION

Figure 4 presents the final FWI VS results for each of the starting models in this study. The
use of different starting models simulates the uncertainty that may exist for the properties of a jet
grout column during a pilot test program. All the inversion results indicate the presence of
anomalous zones within the interior of the overlapping jet grout columns and on the exterior of
each column. These anomalous zones are coincident with the stiffer soil layer as expected.
However, the outline of the true model geometry in Figure 4 clearly reveals that the geometry of
the untreated soil zone is better characterized in the VS results from the 0% initial model. The
overall estimated area is similar to the true model and the geometry of the grouted column
exteriors closely matches the true model. However, the geometry of the elliptical anomalous
zone in the center is grossly overestimated in the horizontal dimension for the 0% initial model
[Figure 4(a)]. The practical implication of this finding is that any adjustments in grouting
parameters (e.g., increase in grouting pressure, etc.) may be excessive if based solely on the
FWI-predicted geometry. In the case of the other initial models, the elliptical anomalous zone in
the center of the two columns has shifted towards the left jet grout column [Figure 4(b) and (c)].
The results from the 20% initial model indicate an intact overlapping zone and instead predicts
that any issues are largely concentrated within the left column. At least some of the anomalous
zone from the 10% initial model coincides with the true model geometry, which represents a
modest improvement relative to the 20% initial model. The same is true for the anomalous region
on the exterior of the two grout columns, with the 20% initial model predicting a greater volume
of intact column than the other results and true model geometry. However, despite better
geometry, the estimated VS for the untreated zone between the two columns is less accurate in the
0% initial model (approximately 260 m/s) than in the 10% and 20% initial models
(approximately 205 m/s), though the difference is minor and does not particularly affect how the
anomaly is interpreted in Figure 4.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 220

The poor results in Figure 4(b) and (c) are likely caused by a local minimum for which the
initial models are in a basin of attraction for the misfit function. The L2-norm function is
nonconvex, which means that local gradient-based optimization cannot guarantee convergence to
the global minimum even with a multi-scale approach (Yao et al. 2019). It is well-documented in
the FWI literature that the starting model plays a large role in the accuracy of the inversion
results due to this issue (Virieux and Operto 2009). Consequently, it is unsurprising that the 0%
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

initial model outperformed the 10% initial model, which outperformed the 20% initial model.
The implication of these findings is that an accurate estimate for the wave propagation properties
of the grout is highly desirable when applying FWI as a diagnostic tool to evaluate the geometry
of untreated zones. This could be accomplished by laboratory testing with bender elements on
grouted soil samples obtained prior to the pilot test program or on coring samples directly
obtained from the column during the pilot test program. The estimated VS from these could then
be used to develop the FWI starting model, and the results in Figure 4(a) highlight that the
predicted geometry of the untreated zone reasonably compares to the true model. Despite all the
discrepancies when grout properties are unavailable, the results in Figure 4(b) and (c) still
identify the presence of an anomalous zone, which is the first step in addressing grouting-related
difficulties at a site.

Figure 4. VS results for different initial models, including column geometry in dashed white
lines: (a) 0% initial model; (b) 10% initial model; and (c) 20% initial model.

CONCLUSIONS

Quality control of jet grout operations continues to present a major challenge. Current
destructive and NDT methods provide limited spatial coverage, resolution, and/or incur
significant costs. The results of this study highlight the potential offered by FWI when applied to
visualize untreated zones in overlapping jet grout columns. The data acquisition efforts were
modest relative to traditional downhole NDT methods since all the sources/receivers were
located at the surface, though the computational efforts are much higher due to the significant
modeling costs. Assuming some a priori information from previous subsurface investigations can
be utilized to estimate the wave propagation parameters for soil/grout, the FWI results
reasonably matched the geometry of the anomalous zone in this numerical study. As less
information was known about the subsurface conditions (specifically the grout properties), the

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 221

FWI results still predicted the presence of untreated zones, but with decreasing accuracy
regarding their spatial extent. It should be noted that this study relied entirely on synthetic data
from numerical simulations, and field application would introduce additional complications
related to noisy data, source characterization, and more complex column geometries. Some of
these field issues could be explored in additional numerical simulations. For example, several
overlapping jet grout columns can be modeled in a three-dimensional random field model to
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

simulate more complex subsurface geometries. However, additional field studies are warranted
to fully evaluate the capabilities of FWI as a diagnostic tool for jet grouting quality control
efforts.

ACKNOWLEDGEMENTS

This research includes calculations carried out on Temple University's HPC resources and
thus was supported in part by the National Science Foundation through major research
instrumentation grant number 1625061 and by the US Army Research Laboratory under contract
number W911NF-16-2-0189. The authors would like to thank Dr. Michael Afanasiev (Mondaic,
Ltd.) for his support with the Salvus software package.

REFERENCES

Afanasiev, M., Boehm, C., van Driel, M., Krischer, L., Rietmann, M., May, D. A., Knepley, M.
G., and Fichtner, A. (2019). “Modular and flexible spectral-element waveform modelling in
two and three dimensions.” Geophysical Journal International, 216(3), 1675-1692.
Bearce, R. G., Mooney, M. A., and Kessouri, P. (2016). “Electrical resistivity imaging of
laboratory soilcrete column geometry.” Journal of Geotechnical and Geoenvironmental
Engineering, 142(3), 04015088.
Bonetto, S., Colombero, C., Comina, C., Giordano, N., Giuliani, A., Mandrone, G., Nicola, S.,
and Tible, P. (2018). “A case study on the application of destructive and non-destructive
methods for evaluating jet-grouting column integrity for bridge-pier scour protection (Cuneo,
NW Italy).” Bulletin of Engineering Geology and the Environment, 77(2), 541–553.
Bunks, C., Saleck, F. Zaleski, S., and Chavent, G. (1995). “Multiscale seismic waveform
inversion.” Geophysics, 60, 1457–1473.
Burke, G. (2012). “The state of the practice of jet grouting.” Grouting and Deep Mixing, 74–88.
Burke, G., and Yoshida, H. (2013). “Jet Grouting.” In K. Kirsch and A. Bell (Eds.), Ground
Improvement, CRC Press, Boca Raton, FL.
Chen, R., Tran, K., La, H., Rawlinson, T., and Dinh, K. (2022). “Detection of delamination and
rebar debonding in concrete structures with ultrasonic SH-waveform tomography.”
Automation in Construction, 104004, 1–33.
Cheng, S. H., Liao, H. J., Yamazaki, J., and Wong, R. K. (2017). “Evaluation of jet grout column
diameters by acoustic monitoring.” Canadian Geotechnical Journal, 54(12), 1781-1789.
Clayton, R., and Engquist, B. (1977). “Absorbing Boundary Conditions for Acoustic and Elastic
Wave Equations.” Bulletin of the Seismological Society of America, 67(6), 1529-1540.
Croce, P., and Modoni, G. (2007). “Design of Jet Grout Cut-off Walls.” Ground Improvement,
11(1), 11-19.
Düzceer, R., and Gökalp, A. (2003). “Construction and Quality Control of Jet Grouting
Applications in Turkey.” Proc. ASCE 3rd International Conference on Grouting and Ground
Treatment, 281–293.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 222

Galindo Guerreros, J. C., Niederleithinger, E., Mackens, S., and Fechner, T. (2015). “Quality
assurance of jet grout columns with borehole seismic measurements.” Proc. International
Symposium Non-destructive Testing in Civil Engineering, 1-7.
Gazzarrini, P. (2021). “A Brief History of Jet Grouting in the Last 50 Years.” GEOSTRATA
Magazine, 25(3), 70–77.
Kordjazi, A., Coe, J. T., and Afanasiev, M. (2021). “Nondestructive Evaluation of Drilled Shaft
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Construction Anomalies Using Full Waveform Tomography of Simulated Crosshole


Measurements.” Journal of Nondestructive Evaluation, 40(1), 1.
Kordjazi, A., Mahvelati, S., Coe, J. T., and Alidoust, P. (2022). “Full Waveform Tomography of
Parallel Seismic Data for Evaluation of Unknown Foundations.” Proc. Geo-Congress 2022,
140–148.
Lee, T. S., Murray, R., and Kiesse, M. (2005). “Jet Grouting at Posey Tube, Oakland,
California.” Proc. Geo-Frontiers 2005, 1–15.
Lin, C. H., Lin, C. P., Ngui, Y. J., Wang, H., Wu, P. L., He, G. J., and Liu, H. C. (2020).
“Diameter assessment of soilcrete column using in-hole electrical resistivity tomography.”
Geotechnique, 70(12), 1120–1132.
Mahvelati, S., and Coe, J. T. (2019). “Comparison of Dispersion-Based Analysis of Surface
Waves and Full Waveform Inversion in Characterizing Unknown Foundations.” Proc. Geo-
Congress 2019, 158–166.
Mooney, M. A., and Bearce, R. (2018). Development of an Electrical Probe for Rapid
Assessment of Ground Improvement. NCHRP IDEA Project 186 Final Report,
Transportation Research Board, The National Academies, Washington, D.C., 39 pp.
Mora, P. (1987). “Nonlinear two-dimensional elastic inversion of multioffset seismic data.”
Geophysics, 52(9), 1211-1228.
Nguyen, T., Tran, T., and McVay, M. (2016). “Evaluation of unknown foundations using
surface-based full waveform tomography.” Journal of Bridge Engineering, 21(5), 04016013.
Nicholson, P. (2014). Soil improvement and ground modification methods. Butterworth-
Heinemann.
Njock, P. G. A., Chen, J., Modoni, G., Arulrajah, A., and Kim, Y. H. (2018). “A review of jet
grouting practice and development.” Arabian Journal of Geosciences, 11(16), 1-31.
Stark, T. D., Axtell, P. J., Lewis, J. R., Dillon, J. C., Empson, W. B., Topi, J. E., and Walberg, F.
C. (2009). “Soil inclusions in jet grout columns.” DFI Journal, 3(1), 33-44.
Tarantola, A. (1984). “Inversion of seismic reflection data in the acoustic approximation.”
Geophysics, 49(8), 1259–1266.
Tromp, J., Komatitsch, D., and Liu, Q. (2008). “Spectral-element and adjoint methods in
seismology.” Communications in Computational Physics, 3(1), 1–32.
Virieux, J., and Operto, S. (2009). “An overview of full-waveform inversion in exploration
geophysics.” Geophysics, 74(6), WCC1-WCC26.
Yao, G., da Silva, N. V., Warner, M., Wu, D., and Yang, C. (2019). “Tackling cycle skipping in
full-waveform inversion with intermediate data.” Geophysics, 84(3), R411-R427.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 223

Surface Wave Site Characterization with MATLAB and Geopsy

Dennis R. Hiltunen, M.ASCE1


1
Professor, Dept. of Civil and Coastal Engineering, Univ. of Florida, Gainesville, FL.
Email: dhilt@ce.ufl.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ABSTRACT

An assortment of algorithms for the processing of surface wave data are presented and
utilized to analyze several case studies. Algorithms developed in MATLAB are described for
extracting dispersion curves from waveforms for all typical field test scenarios, including SASW,
MASW, ReMi, and 2D passive arrays. Geopsy contains a publicly available module that inverts
dispersion curves for both shear wave velocity and layer thickness based upon a global search
algorithm that provides valuable information on the uncertainty of the profile parameters. A suite
of models from Geopsy can be post processed with MATLAB routines for display of inversion
results, including important statistical parameters. The algorithms have been used to process
various forms of original data from several case studies. Inversions with lowest variability are
produced by soft soil sites with gentle variation, whereas sites with sharp contrasts such as soil
over rock at shallow depths can produce inversion results with much larger variability. Inversion
for that “one true model” can be elusive, even in the case of synthetic data sets where the true
model is known; it is possible to come close, but with inversion come both uncertainty and
blurriness.

INTRODUCTION

Determination of a shear wave velocity profile via in situ surface wave methods requires
three fundamental steps (Figure 1): 1) collection of vibration waveforms via a field experiment,
2) extraction of the surface wave dispersion relationship (e.g., phase velocity versus frequency)
from the waveforms, and 3) inversion of the dispersion curve.
Algorithms available from the author and developed in MATLAB are described for
extracting dispersion curves from waveforms for all typical field test scenarios, including,
SASW, MASW, ReMi, and 2D passive arrays. Geopsy (Wathelet et al. 2020) contains a publicly
available module that inverts dispersion curves for both shear wave velocity and layer thickness
based upon a global search algorithm rather than traditional linearized techniques, which
provides valuable information on the uncertainty of the profile parameters. A MATLAB post-
processing module has been developed for displaying in both 2D and 3D the results from
Geopsy, including the inversion profile and fit of the experimental data, but also important
statistics that quantify the variability in the profiles and fit. Results from the algorithms are
displayed and described via several case studies.

MATLAB DISPERSION ANALYSIS

Algorithms developed in MATLAB are available for extracting dispersion curves from field
waveforms for all typical field test scenarios, including:
• Spectral analysis of surface waves (SASW)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 224

• Linear, multiple sensor, active source arrays with both impact and shaker sources (e.g.,
MASW)
• Linear, multiple sensor, passive source arrays (e.g., ReMi)
• Passive arrays employing two dimensional arrangements of surface sensors.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Surface Wave Steps (from Foti et al. 2018)

SASW: The SASW techniques introduced by Stokoe, Nazarian and others in the 1980s
(Stokoe et al. 1994) made use of Fourier transform algorithms to decompose the signals
measured with two vibration transducers and generated with multi-frequency impact sources or
vibration shakers. It is typical that a dynamic signal analyzer is utilized for collection of field
records and to observe the frequency domain cross power spectrum phase angle and coherence
functions during testing to ensure that quality records are collected for further processing.
MATLAB algorithms accept these frequency domain functions and then process the records for
determination of the dispersion relationship. The processing tasks include masking to remove
poor quality data, phase angle unwrapping, calculation of the phase velocity for each frequency,
and assembling of a combined dispersion curve from records collected at multiple receiver
spacings. If desired, smoothing of the combined dispersion curve can be conducted prior to
inversion.
MASW: For the multiple sensor, active arrays, several 2D transform techniques can be
utilized, including frequency-wavenumber (f-k), frequency-slowness (f-p), Park et al. transform,
traditional beamformer, and cylindrical beamformer. Tran (2007) provides a detailed description
of the methods. The f-k and f-p transforms, the Park et al. (1998) transform, and traditional
beamformers all treat the multi-channel record as a plane wavefield, while a cylindrical
beamformer (Zywicki 1999) uses the cylindrical wave equations to transform and identify
Rayleigh waves. The cylindrical wave equations describe active surface wavefields more
properly than do plane wave equations, since the source is a finite distance from the receivers.
This allows estimating phase velocities for relatively long wavelengths compared to the array
length of receivers (Zywicki 1999). MATLAB algorithms accept time domain records collected
in the field at multiple receiver locations (x-t domain) and transform to the phase velocity-
frequency domain. The dispersion relationship is then determined by picking the peaks of the 2D
transform, either automatically or via the MASWaves routine described by Olafsdottir et al.
(2018). The MASWaves routine can also combine dispersion relationships from multiple source
locations, receiver spacings, etc., into a single relationship for inversion.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 225

In addition, techniques have emerged to further improve the accuracy and resolution of
wavefield characterization, particularly at low frequencies, which serves to increase the depth of
characterization. Using a small vibration shaker, a linear array of non-uniformly spaced
geophones, and array-based cylindrical beamformer two-dimensional wavefield transformations,
Hebler (2001) and Jiang et al. (2015) demonstrated significantly improved determination of the
Rayleigh wave dispersion relationship. The MATLAB routines will accept shaker-based
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

waveforms typically collected in the frequency domain to determine the dispersion relationship
using the 2D transform, picking, and combining routines discussed above.
ReMi: The coupling of a linear, multi-channel array, 2D wavefield transformations, and so-
called passive vibration sources available in the ambient environment (e.g., vehicle traffic,
industrial machinery, movement of water) produced the now well-known ReMi techniques first
presented by Louie (2001). Since the passive vibration sources typically provide increased
energy at low frequencies than available from a sledgehammer active source, these passive
techniques have typically increased the depth of characterization at most sites without the
requirement of a large impact source or specialized vibration shaker. MATLAB algorithms
developed and described in detail by Li (2008) will transform ambient noise records collected in
the field (x-t domain) to the phase velocity versus frequency domain for determination of the
dispersion relationship. Multiple noise records can be stacked to improve the quality of the
image and a picking routine is available for implementing the recommendations of Louie in
producing the phase velocity value at each frequency.
2D Passive: Several research studies have utilized arrays of transducers arranged two-
dimensionally (e.g., circle, triangle, L-shape) to collect wavefield signals from passive vibration
sources. Since the location of a passive source is typically unknown, a 2D array can resolve both
the speed and the direction of wave propagation, which provides a more accurate determination
of wavefield velocity than can typically be accomplished with a linear array (Rosenblad and Li
2009). MATLAB algorithms are available for processing records from 2D passive arrays,
including, Capon's MVDL, MUSIC, and frequency domain beamformer and which are well
described by Zwicki (1999). The frequency domain beamformer (FDBF) for 2D arrays was
proposed by Lacoss et al. (1969) and is a direct extension of the beamformer utilized for 1D
linear arrays. Capon (1969) introduced a high-resolution beamforming method called minimum
variance distortionless look (MVDL) to reduce the effects of signal contributions from directions
of arrival other than along the active look direction. Multiple signal classification (MUSIC) is
another high-resolution beamforming method proposed by Schmidt (1986) whose power
estimate is like Capon’s MVDL. The MATLAB algorithms will transform ambient noise records
collected in the field (x-t domain) to the phase velocity versus frequency domain for
determination of the dispersion relationship. Multiple records can be stacked to improve the
quality of the image and the phase velocity value at each frequency is automatically picked from
the transformation image.
Forward Models: Calculation of a theoretical dispersion relationship from a stack of
homogeneous layers is a primary component of any surface wave inversion methodology.
Several algorithms are publicly available. The Geopsy suite of programs described below has a
very fast implementation of the Thompson (1950) and Haskell (1953) technique coded and
compiled with C++ and available as a separate module. Marosi and Hiltunen (2001) describe an
implementation of Knopoff’s method coded and compiled in Fortran. Lai (1998) developed the
DISPERSE routine and Olafsdottir et al. (2018) the method of Kausel and Roessett (1981), both
in MATLAB.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 226

GEOPSY INVERSION

Inversion of the dispersion relationship to determine a shear wave velocity profile has
traditionally been conducted with local search algorithms employing linearized least-squares
techniques. Stochastic, global search algorithms have been implemented more recently as the
speed of computing has increased. For either option, it has been typical to invert for the unknown
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

shear wave velocity of a stack of homogeneous, horizontal layers of fixed thickness and assumed
Poisson’s ratio and mass density.
The DINVER module of the software suite Geopsy was introduced by Wathelet et al. (2020)
and is available in the public domain. DINVER utilizes a global search methodology known as
the neighborhood algorithm and inverts for unknown values of both shear wave velocity and
layer thickness, a significant improvement over all previous methods in the experience and
opinion of this author. For the forward model, DINVER utilizes a very fast implementation of
the Thompson and Haskell methodology for determination of the dispersion relationship for each
sampled model of the shear wave velocity profile.
The success of local search techniques is often dependent on the selection of a close to actual
starting model that includes assumption of the thickness for each of the stacked layers. To start,
DINVER requires specification of a range in shear wave velocity and thickness values for each
layer and successful inversion can be accomplished with wide ranges indicated for the inputs.
Inversions conducted with local search algorithms typically provide one shear wave velocity
profile that has been found to best fit the experimental dispersion data for the given stack of
layers. Like most global techniques, DINVER provides a suite of models that fit the data within a
range specified by the user. This suite can provide not only the best fit model, but also a sample
of models from which statistics can be calculated and observed to assess uncertainty parameters
of the shear wave velocity profile. A graphics module part of DINVER provides visual
observation of the suite of shear wave velocity models for a user specified degree of fit, an
example of which is shown in Figure 2. In part a, the target dispersion curve is shown with
circular solid symbols and each point includes error bars of one standard deviation to
approximate measurement error. The dispersion curves sampled by the inversion process and
within approximately three standard deviations of the model that best fits the target are shown in
a shaded histogram format using a range of colors, with those in gray being most abundant and
those in dark red having lowest frequency of occurrence. The shear wave velocity profiles
responsible for the dispersion curves in part a are shown in a similar histogram format in part b.

MATLAB POST PROCESSING

As discussed in Figure 2, DINVER provides a graphical image of the shear wave velocity
profiles and dispersion curve fits for the suite of output models. It is also possible to export the
suite of models for further processing. The images from DINVER are helpful, but they can be
somewhat misleading regarding the extent of variability, since they show all the dispersion
curves via color coding. Thus, a MATLAB routine has been developed to provide further
information. Common statistical parameters of histogram, mean, mode, standard deviation,
coefficient of variation, and error bands are computed from the exported suite of models and
dispersion curves and can be plotted jointly on both 2D and 3D images along with the best fit
model and dispersion relationship. By way of example, Figure 3 presents 2D images for the same
data set previously shown in Figure 2. Here, rather than all the dispersion curves, a histogram is

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 227

shown in shades of blue, the best fit model is shown in red, the mean and mode of all models is
shown via asterisk data points, and the one standard deviation (SD) range of all models is shown
via cyan lines. The replotting using the computed histograms and statistics provides a more
realistic visual depiction of the variability, as the color shading technique in Figure 2 tends to
exaggerate the actual variability. As noted in Figure 3, the variability in both dispersion data and
shear wave velocity is small for the BM10 synthetic example. As well, statistical parameters can
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

be reported to characterize variability of a shear wave velocity profile. Since the example shown
is for the BM10 synthetic, the true shear wave velocity profile is shown via green line. It is noted
that, while the best fit model from the inversion is similar, it does not completely recover the
known true model but a close approximation.

(a) Dispersion Curves (b) Shear Wave Velocity Models

Figure 2. DINVER Graphics Output from Vantassel and Cox (2021) BM10 Analysis Using
Three Standard Deviations from Best Fit Model

(a) Dispersion Curves (b) Shear Wave Velocity Models

Figure 3. MATLAB Post Processing of Vantassel and Cox (2021) BM10 Using Three
Standard Deviations from Best Fit Model

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 228

CASE STUDIES

The MATLAB algorithms and DINVER module have been used to process various forms of
original data from several case studies summarized below. Emphasis is placed on displaying and
discussing the uncertainty results from the DINVER inversions of all cases and example results
are shown herein. The case studies included six synthetics, seven soil sites, and six rock sites.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Synthetics: Six case studies were inverted with the DINVER module using synthetic
dispersion relationships, three typical of geotechnical sites (up to 30 m depth) and three deep
profiles (greater than 100 m), and each set of three contains a range of stiffness variation. These
synthetics provide a useful introduction to learning the DINVER module. The dispersion
relationships for the three geotechnical sites were produced by Vantassel and Cox (2021) and are
displayed in Figure 4a as BM10, BM11, and BM12. The dispersion relationships for the three
deep profiles were produced by Li (2008) and are displayed in Figure 4b as Case 1, Case 2, and
Case 3.
Figures 2 and 3 presented the results for the BM10 synthetic. This profile has a gentle
increase in stiffness with depth; the DINVER inversion process worked well for this profile and
the variability in profiles for a range of dispersion curves of about three standard deviations from
the best-fit model is relatively small. In contrast, the results for BM11 in Figure 5 display a much
larger variation in the shear wave velocity profiles even though the variability in dispersion data
is small and like Figure 3a. As noted, BM11 contains sharper increases in stiffness at shallow
depth. Again, for BM11, it is noted that even though the true synthetic model is known, the
inversion process does not exactly find the true model, but a close approximation.

(a) Geotechnical Sites (up to 30 m) (b) Deep Sites (>100 m)

Figure 4. Dispersion Curves for Synthetic Case Studies

Soil Sites: Seven cases come from field test data collected over a wide range of profile
conditions, from soft clay to gravel deposits. The original data were collected in various
configurations of SASW, multi-sensor, and active and passive field techniques and the dispersion
relationships are shown in Figure 6a. The four softest sites (CL, HLP, P, and TAMU) with clays,
silts, and/or sands in the material and that have a more gradual increase in phase velocity with
decreasing frequency, all inverted well and produced small variation in the profiles. By way of

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 229

example, Figures 7a and 7b display the inversion results for the TAMU site. In contrast, for the
remaining three soil sites with coarse-grained soils and gravels in the profile (LaS, M, ST) and
with higher phase velocity and a larger increase in phase velocity with decreasing frequency, the
inversion worked well, but the variability in the inverted profiles is much larger even though the
variability in dispersion data is small. By way of example, Figures 7c and 7d display the
inversion results for the LaS site. A final item to note is that all seven profiles were successfully
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

inverted using a DINVER model that specified a maximum depth for the inversion equal to one-
half of the longest wavelength available from the dispersion relationship.

(a) Dispersion Curves (b) Shear Wave Velocity Models

Figure 5. MATLAB Post Processing of Vantassel and Cox (2021) BM11 Using Three
Standard Deviations from Best Fit Model

(a) Soil Sites (b) Rock Sites

Figure 6. Dispersion Curves for Soil and Rock Field Test Sites

Rock Sites: Finally, six case studies were from field test data collected at sites with rock at
shallow depths. Again, the original data were collected in various configurations of SASW,
multi-sensor, and active and passive field techniques and the dispersion relationships are shown

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 230

in Figure 6b. All six rock sites were successfully inverted, and two sets of results are shown in
Figure 8. In contrast to the soil sites, it was found that more care is required in inverting for the
profiles with DINVER to ensure that the sampled models produced are not wildly variable. It
was typically required that the maximum depth specified in the inversion model could not exceed
one-third of the maximum wavelength, as opposed to one-half for the soil sites. Even then, the
inverted models are only reasonable to a depth of the largest stiffness contrast attributable to the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

rock. Below the large contrast, the variation in shear wave velocity values is large even though
the variability in dispersion data is small; reliable quantification of the stiffness of the rock
material is questionable. The one exception is the PSU site with inversion results presented in
Figures 8a and 8b; here the variation in models is reasonable. The exception is due to the “S-
shaped” dispersion relationship as opposed to the “angle-shaped” relationships for the other five
sites. A large increase in phase velocity with decreasing frequency such as for the Newberry site
produces a poorly resolved inversion at depth. The PSU dispersion relationship with a flattening
of the dispersion curve at the lowest frequencies allows the inversion to reliably resolve the shear
wave velocity of the deepest layers.

(a) Dispersion Curves for CL (b) Shear Wave Velocity Models for CL

(c) Dispersion Curves for LaS (d) Shear Wave Velocity Models for LaS

Figure 7. MATLAB Post Processing of Two Soil Field Sites Using Three Standard
Deviations from Best Fit Model

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 231
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(a) Dispersion Curves for PSU (b) Shear Wave Velocity Models for PSU

(d) Shear Wave Velocity Models for


(c) Dispersion Curves for Newberry
Newberry

Figure 8. MATLAB Post Processing of Two Rock Field Sites Using Three Standard
Deviations from Best Fit Model

CONCLUSIONS

An assortment of algorithms for the processing of surface wave test data have been presented
and utilized to analyze several case studies. A summary of findings is presented as follows:
• MATLAB routines are available for processing test data into dispersion relationships for
all typical field test implementations.
• The Geopsy DINVER module is a superior methodology for surface wave inversion due
to its ability to invert for both shear wave velocity and layer thickness and to produce
multiple models for assessment of variability.
• A suite of models from DINVER can be post processed with MATLAB routines for
display of inversion results including important statistical parameters.
• Inversions with lowest variability are produced by soft soil sites with shear wave velocity
less than about 300 m/s and whose stiffness increase with depth is gentle.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 232

• As both the magnitude and gradient of the stiffness increases, the variability in the
inverted profiles also increases.
• The shallow rock sites with “angled-shape” dispersion curves are the worst, producing
large variability at depth and shear wave velocity models that are of questionable
reliability; the best-fit model may be usable, but the variation is large.
• The “S-shaped” dispersion results (TAMU and PSU) produce models with reasonable
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

variability even in the presence of stiffness contrasts at depth, which further illuminates
the importance of collecting reliable surface wave data at low frequencies.
• The inversion for that “one true model” is elusive, even in the case of synthetic data sets
where the true model is known; it is possible to come close, but with inversion comes
both uncertainty and blurriness.

REFERENCES

Capon, J. (1969). “High-Resolution Frequency-Wavenumber Spectrum Analysis,” Proceedings


of the IEEE, Vol. 57, No. 8, pp. 1408-1418.
Foti, S., et al. (2018). “Guidelines for the Good Practice of Surface Wave Analysis: A Product of
the InterPACIFIC Project,” Bulletin of Earthquake Engineering, Vol. 16, pp. 2367-2420.
Haskell, N. A. (1953). “The Dispersion of Surface Waves on Multilayered Media,” Bulletin of
the Seismological Society of America, Vol. 43, No. 1, January, pp. 17-34.
Hebeler, G. L. (2001). Site Characterization in Shelby County, Tennessee Using Advanced
Surface Wave Methods, M.S. Thesis, Georgia Institute of Technology.
Jiang, P., Tran, K. T., Hiltunen, D. R., and Hudyma, N. (2015). “An Appraisal of a New
Generation of Surface Wave Techniques at a Test Site in Florida,” IFCEE 2015,
Geotechnical Special Publication No. 256, M. Iskander, M. T. Suleiman, J. B. Anderson, and
D. Laefer, Eds., American Society of Civil Engineers, pp. 1981-1992.
Kausel, E., and Roesset, J. M. (1981). “Stiffness Matrices for Layered Soils,” Bulletin of the
Seismological Society of America, Vol. 71, No. 6, pp. 1743–1761.
Lacoss, R. T., Kelly, E. J., and Toksoz, M. N. (1969). “Estimation of Seismic Noise Structure
using Arrays,” Geophysics, Vol. 34, pp. 21-38.
Lai, C. G. (1998). Simultaneous Inversion of Rayleigh Phase Velocity and Attenuation for Near-
Surface Site Characterization, Ph.D. Dissertation, Georgia Institute of Technology.
Li, J. (2008). Study of Surface Wave Methods for Deep Shear Wave Velocity Profiling Applied to
the Deep Sediments of the Mississippi Embayment, Ph.D. Dissertation, University of
Missouri.
Louie, J. N. (2001). “Faster, Better, Shear-Wave Velocity to 100 Meters Depth from Refraction
Microtremor Arrays,” Bulletin of Seismological Society of America, Vol. 91, No. 2, pp. 347-
364.
Marosi, K. E., and Hiltunen, D. R. (2001). “Systematic Protocol for SASW Inversion”
Proceedings of the Fourth International Conference on Recent Advances in Geotechnical
Earthquake Engineering and Soil Dynamics, San Diego, March 26-31.
McMechan, G. A., and Yedlin, M. J. (1981). “Analysis of Dispersive Waves by Wave Field
Transformation,” Geophysics, Vol. 46, No. 6, pp. 869-871.
Olafsdottir, E. A., Erlingsson, S., and Bessason, B. (2018). “Tool for Analysis of Multichannel
Analysis of Surface Waves (MASW) Field Data and Evaluation of Shear Wave Velocity
Profiles of Soils,” Canadian Geotechnical Journal, Vol. 55, pp. 217-233.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 233

Park, C. B., Xia, J., and Miller, R. D. (1998). “Imaging Dispersion Curves of Surface Waves on
Multi-Channel Record,” Expanded Abstracts, 68th Annual Meeting of Society of Exploration
Geophysicists, pp. 1377-1380.
Rosenblad, B. L., and Li, J. (2009). “Performance of Active and Passive Methods for Measuring
Low-Frequency Surface Wave Dispersion Curves, Journal of Geotechnical and
Geoenvironmental Engineering, American Society of Civil Engineers, Vol. 135, No. 10,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

October, pp. 1419-1428.


Schmidt, R. O. (1986). “Multiple Source DF Signal Processing: An Experimental System,” IEEE
Trans. Ant. Prop, Vol. AP-34, No. 3, pp. 281-290.
Stokoe, K. H., II, Wright, S. G., Bay, J. A., and Roesset, J. M. (1994). “Characterization of
Geotechnical Sites by SASW Method,” Geophysical Characterization of Sites, R. D. Woods,
Ed., Oxford & IBH Publishing Co., New Delhi, India, pp. 15-25.
Thomson, W. T. (1950). “Transmission of Elastic Waves Through a Stratified Soil
Medium,”Journal of Applied Physics, Vol. 21, No. 2, February, pp. 89-93.
Tran, K. T. (2007). An Appraisal of Surface Wave Methods for Soil Characterization, M.Eng.
Thesis, University of Florida, 77 pp.
Vantassel, J. P., and Cox, B. R. (2021). “SWinvert: A Workflow for Performing Rigorous 1-D
Surface Wave Inversions,” Geophysical Journal International, Vol. 224, pp. 1141-1156.
Wathelet, M., et al. (2020). “Geopsy: A User-Friendly Open-Source Tool Set for Ambient
Vibration Analysis,” Seismological Research Letters, Vol. 91, No. 3, May, pp. 1878-1889.
Zywicki, D. J. (1999). Advanced Signal Processing Methods Applied to Engineering Analysis of
Seismic Surface Waves, Ph.D. Thesis, Georgia Institute of Technology, 357 pp.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 234

Geo-Acoustic Signals in Geotechnical and Foundation Engineering

Anisha Pokhrel, S.M.ASCE1; and Sherif L. Abdelaziz, Ph.D., A.M.ASCE2


1
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
VA. Email: anishap@vt.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
VA. Email: saziz@vt.edu

ABSTRACT

The goal of this study is to summarize the existing state of practice for using geo-acoustic
signals in geotechnical and foundation engineering and to further outline potential future
directions for this technique. For decades, acoustic signals have been recognized as a feasible
geophysical technique to estimate soil properties including, but not limited to, the shear wave
velocity in a soil medium, porosity, and relative density. These geophysical acoustic methods
rely on signals produced by an external source within the soil mass. Recent studies further
suggested potential correlations between acoustic signals produced naturally by soil formations
and the physical and mechanical properties of these soils. In this manuscript, we will first
summarize current knowledge about the current uses of geo-acoustic signals from the literature.
We will then present a brief discussion about existing applications of such technology in
geotechnical and foundation engineering.

INTRODUCTION

Acoustic signals have been regarded as a viable geophysical tool for estimating soil
parameters such as shear wave velocity in a soil medium, porosity, and relative density. The
current practice of the acoustic signal relies on the signals produced from the external sources.
However, various research studies suggest that it is possible to detect the acoustic waves
generated naturally from the soil due to shear deformation. Using acoustic real-time monitoring,
it is found that there is a direct relationship between the deformation rate and acoustic signal
emission rate (Dixon et al., 2015). Furthermore, acoustic emission monitoring has been used to
quantify the accelerating deformation behavior in geotechnical systems by categorizing them
into four groups (rapid, moderate, slow, very slow) based on slope deformation rate and acoustic
emission event rate, allowing for early detection of landslides (Dixon and Spriggs, 2007). Recent
studies further suggested potential correlations between acoustic signals produced naturally by
soil formations and the physical and mechanical properties of these soils, for instance, effect of
water content, effect of different stress paths, and disturbance effect of wave guides. This paper
summarizes the current knowledge regarding uses of acoustic signals in the field of geotechnical
engineering.

GENERATION OF ACOUSTIC SIGNALS

Acoustic waves are generated from soils due to different deformation phenomena like
friction, particle arrangement, crushing, rolling of the particles and so on. Our main aim is to
relate these generated acoustic signals with the soil deformation under local straining in real-time

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 235

with the help of various devices like transducers, sensors, geophones, etc. Generally, geological
materials emit a wide range of acoustic signal frequency from 10Hz to 500kHz (Dixon et al.,
2003). The high frequency acoustic signals from 30kHz to 80 kHz appear to be directly linked to
continual grain-scale interactions such as friction and rolling (Michlmayr et al., 2013). Particle
microcracking and crushing mechanisms generate higher frequency acoustic emission >100kHz
than particle interaction mechanisms such as friction and rearrangement (<100kHz) (Lin et al.,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2019). However, background noise can be problematic to the field research generating
extraneous signals. Typically, the frequency of noise due to vehicles and activities of the general
public is below 15 Hz (Dixon et al., 2003). It is necessary to use noise cancellations or filters
with the sensors (like geophones, accelerometers, acoustic emission (AE) transducers,
hydrophones) to improve the results by minimizing the potential for false alarms. Although high
frequencies are less affected by background noises, the corresponding signals are more affected
by the attenuation of the geological materials like soils and rocks. Soils have relatively high
attenuation of the AE signals, which could loss the information about the signals > 100kHz
(Dixon et al., 2003). Therefore, it necessitates the use of waveguides having low attenuation of
the signals (e.g., steel tubes) through which the generated acoustic emission propagates and the
signals are detected by the sensor attached to the top of the waveguide.

USE OF WAVEGUIDE AND WAVE GENERATOR

Waveguides and wave generator (backfill) are important elements in real-time monitoring
system of acoustic signals and have been used in various research efforts (Dixon et al., 2003;
Caicedo and Patino-Ramirez, 2019; Dixon and Spriggs, 2007; Dixon et al., 2015; Dixon et al.,
2015; Codeglia et al., 2017). The AEs generated due to soil deformation are dependent on the
nature of the soil. It is reported that AE generated in each section of a rock slope is different and
transmits differently through the given geological conditions (Cheon et al., 2011). Soil has high
attenuation of the signals, which results in poor conveyance of the produced signals to the
sensors and necessitates waveguides. The use of waveguide in slope monitoring is shown in Fig
1.
These waveguides help in proper conveyance of acoustic signals from the source to acoustic
sensor. These acoustic signals are produced by deformation of the backfill to the waveguide for
instance, the steel tube, in response to deformation of the host soil. Since, it has a significant role
in detecting acoustic waves, correct waveguide material (having less attenuation) should be
selected, otherwise it may display less accurate results. The relationship between AE and
deformation is dependent on the backfill material and waveguide, which can affect the threshold
level for triggering a warning. The AE generated from the soil is also dependent on the nature of
the backfill material (i.e., wave generator). Different materials have different attenuation due to
mechanisms such as geometrical spreading, internal friction, scattering, diffraction and
dispersion (Smith and Dixon, 2020). This creates an issue regarding poor quantification of
acoustic signals that may result fake signals/alarms.
Multiple studies have explored waveguides and generators. Dixon et al. (2003) explored the
effects of different backfill such as sand, gravel and Bentonite grout and found that sand backfills
gave a better response than others because sand can remain unstable for a longer period
following a de-stabilization event (Dixon et al., 2003). Dixon et al. (2003) also confirms that
coarse-grained soil backfills produce the highest levels of AE, while the fine-grained soils
produce the lowest.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 236

Deng et al. (2019) studied four different type of waveguides and backfills and showed that
the combination of the copper tube and backfill material as the active waveguide exhibited
remarkable variation characteristics and strong linear correlation between AE and the
deformation parameter (Deng et al., 2019). Smith et al. (2017) investigated AE attenuation due to
soil coupling and found that the use of river gravel as the backfill material allowed the
propagation of the stress waves over 10m at the lowest threshold level whereas clay materials
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

resulted in severe damping of the signal to below the detection threshold by propagation of less
than 4 m from the source. However, there is an issue regarding the source of the emitted wave
signals, which can either be wave-generator or the soil or rock-mass. Also, acoustic waves can
get reflected back along the waveguides therefore, the coefficient of reflection could impact the
propagation of the AE along the waveguide. Smith et al. (2017) suggested that the attenuation
coefficient is dependent on reflection coefficient.

RELATIONSHIP BETWEEN ACOUSTIC EMISSION RATE AND SOIL


DEFORMATION

Various studies have shown that a direct correlation exists between the acoustic emission rate
and soil deformation rate (Dixon and Spriggs (2007); Pioro and Oshkin (2011); Dixon et al.
(2015); Smith and Dixon (2020)). Dixon and Spriggs (2007) shows the relationship between
acoustic event rate and displacement rate to be directly proportional to each other over a range
that is consistent with standard landslide movement classification (1-0.001 mm/min) from which
both increase and decrease in deformation rate can be identified. The schematic of slope AE
monitoring system used in (Dixon and Spriggs, 2007) is shown in Fig. 1. It was also found that
AE is proportional to mobilized shear strength, and can be used to quantify the accelerating
deformation behavior in geotechnical systems (Smith and Dixon, 2020). As stated earlier, AE
rate is found to be proportional to displacement rate as shown in Fig. 2. In addition to that, the
measured AE energy was proportional to mobilized penetration resistance in case of pile
foundation and AE approach detects accelerating deformation behavior of a pipe/soil system
(Smith and Dixon, 2020).

Fig. 1. Schematic of slope AE monitoring system (Dixon and Spriggs, 2007).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 237
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Relationship between AE rate and imposed displacement rate (Caicedo and Patino-
Ramirez, 2019).

Ring-down Count (RDC) per hour is one of the effective parameters to measure the acoustic
emission rate. RDC measures a count of signal threshold crossings, with the threshold value
predetermined to eliminate background and/or electronic noise. Deng et al. (2019) presents the
linear relationships between different parameters, such as the RDC and velocity, the cumulative
RDC and displacement, and the RDC and thrust. Although the rate of change of AE maintained
the same trend as acceleration, the relationship between AE and deformation was related to the
backfill material and waveguide material, which affected the threshold level for triggering a
warning (Deng et al., 2019). There are still numerous obstacles to developing broadly applicable
AE interpretation methodologies that can be used by decision makers to convey information on
the status of a variety of geotechnical assets and hence provide early warning of system
breakdown.

QUANTIFICATION OF SLOPE FAILURE

The quantification of slope failure is done by categorizing the mode of slope instability into
four different groups: rapid, moderate, slow and very slow on the basis of slope deformation rate
(mm/min) and gradient of AE event rate (event counts/hr2) (Dixon and Spriggs, 2007), as shown
in Table 1. However, Koerner et al. (1978) developed a different quantification of slope failure
on the basis of AE rates stating that soil masses that do not generate acoustic emissions are
probably not deforming and are therefore stable. Meanwhile, soil masses that generate moderate
levels of acoustic emissions (from 10 counts/min to 100 counts/ min for the soil types, equipment
and sensitivities used herein) are deforming slightly and are to be considered marginally stable.
Moreover, soil masses that generate high levels of acoustic emissions (from 100 counts/min to
500 counts/min for the soil types, equipment and sensitivities used herein) are deforming
substantially and are to be considered unstable. Finally, soil masses that generate very high levels
of acoustic emissions (greater than 500 counts/min for the soil types, equipment and sensitivities

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 238

used herein) are undergoing large deformations and can be considered to be in a failure state.
This shows that there might be some discrepancies in the quantification.

Table 1. Assessment of slope displacement rates and state using quantified AE generated
by the active waveguide used in (Dixon and Spriggs, 2007).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Slope deformation Gradient of AE State of slope


rate (mm/min) event rate
(events/min2)
Rapid >1 >100 Slope is undergoing large
deformation and is likely to be
in a state of failure. Urgent need
to implement public safety
measures
Moderate 0.1 100 to 1 Substantial deformations and
considered unstable, immediate
remedial and public safety
measures required
Slow 0.01 1 to 0.01 Deforming slightly but
marginally stable, continued
monitoring is necessary
Very <0.001 <0.01 Slope can be considered
Slow essentially stable

(Dixon et al., 2015) confirms that the sensitivity of AE to small displacements through
comparisons with continuous ShapeAccelArray (SAA) deformation measurements during a
movements event of 0.0075 mm. Moreover, this study demonstrated the direct correlation
between AE rates generated by the active waveguides and the velocity of slope movement,
which can be used to detect changes in rates of movements, such as accelerations and
decelerations in response to both destabilizing and stabilizing effects (Dixon, Smith, et al., 2015).

PATTERNS OF ACOUSTIC EMISSION

There are various mechanisms that cause the deformation of soils, such as variation in
groundwater table, snow loading, earthquakes, etc. Codeglia et al. (2017) categorizes acoustic
emission patterns into three different groups as per the order of counts/hour: Type A, Type B and
Type C with hundreds counts/hour, tens of thousands count/hour and hundreds of thousands
counts/hour, respectively (Fig. 3). Different ranges of acoustic emissions are associated with
different stabilizing or destabilizing factors of soil like rainfall, infiltration, and snow loading.
From the trial site of Passo della Morte, it was hypothesizeds that type A events are generated by
local mechanisms such as deformation on a discontinuity or local groundwater flow from which
the low energy AEs are generated that cannot propagate to more than one waveguide (Codeglia
et al., 2017). In the same way, for Type B, AEs are associated with changes in groundwater level
some of which do not correlate to particular rainfall event, thus, the mechanism is unclear
(Codeglia et al., 2017). Type C AEs are predicted to be generated by snow loading on the surface
of the slope (Codeglia et al., 2017). Therefore, it can be concluded that AE generated due to

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 239

different mechanisms can be differentiated based upon the range of acoustic emission rate.
However, the correlation between rainfall and AE rates are not consistent according to the site
experiment. Furthermore, the analysis of this experiment clearly concludes that there is no
acoustic emission response associated to recorded earthquakes (Codeglia et al., 2017).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. Type A, Type B, Type C acoustic emission patterns at Passo della Morte (Codeglia et
al., 2017).

EXAMPLES APPLICATION OF GEO-ACOUSTIC SIGNALS

Slope Stability. Since the detection of acoustic signals generated because of ongoing
deformation in the slope is possible and the generated AE rates are proved to be proportional to
the rate of slope movement (Berg et al. (2018); Dixon and Spriggs, 2007; Dixon, Spriggs, et al.
(2015); Dixon, Smith, et al. (2015)), developing early warning systems for slope instabilities has
great potential. With proper quantification and categorization of the detected signals, alert
messages according to the state of the signals can be conveyed to the public and thus early
precaution measures could be implemented as per the state of the anticipated events.
Foundation Engineering. Geo-acoustic signals can also be used in foundation engineering
such as for construction of piles. Mao et al. (2018) performed pile penetration test with two
samples: a silica sand and a coral sand. Figure 4 shows the schematic diagram of the instrument
set up for acoustic signals detection. Generally, it was found that the AE originating from the
impending fracture failure of the particles was accompanied by a significant rise in high
frequency components (>100 kHz) (Mao et al., 2018). The mobilized penetration resistance was
found to be directly related to measured acoustic emission energy (Mao et al., 2018). Also, AE
characteristics can also be used to distinguish the subsoil response of the highly stressed soils
below the pile tip (Mao et al., 2016).

POTENTIAL AND CHALLENGES OF AE

Geotechnical infrastructure systems are interconnected to many other disciplines such as


transportation, buildings, and hydropower systems. Thus, these infrastructure systems are related
to many aspects of public property and lives and it is important to develop a mechanism to
provide for their efficient maintenance. For this, the development of early warning systems could
play a crucial role in enhancing quality of life and systems. According to various studies,
acoustic signals generated from on-going soil deformations have huge potential in the design of
early warning systems. However, there are still some discrepancies among the parameters of the

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 240

acoustic signal and physical properties of soils. The acoustic emissions in soils are generated not
only because of the soil deformation but also can be because of natural activities like rainfall,
infiltration (Cheon et al., 2011), movement of groundwater level, and frost heave action (Berg et
al., 2018). The induced acoustic emission due to these mentioned causes may super-impose with
the real acoustic signals generated due to soil deformation. This can result in issues regarding the
correlation between the damage-related acoustic emission in the waveguide assembly and the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

instability in the slope. Cheon et al. (2011) clearly posits that acoustic emission rates can vary
with slope displacement rates and water pressure in unstable slopes, although rock slopes have
small displacement until their failure and the water pressure is very small compared to the stress
induced by the displacements. Berg et al. (2018) shows that the extraneous noise detected due to
thermal effect results in significant discrepancy between the obtained result and the
quantification done in the framework developed by Dixon, Spriggs, et al. (2015). Hence, the
issues regarding factors affecting the acoustic emissions should be resolved, otherwise the results
might be affected and standard criteria for stability evaluation might not be strong enough to
determine the state of slope.

Fig. 4. Schematic diagram of pile test system equipped with AE monitoring (Mao et al.,
2018).

Furthermore, the acoustic signals generated from the shearing of soils are also dependent on
the properties of the soil for instance, shape, size, moisture content and plasticity. Laboratory test
for both isostatic and triaxial creep mode under drained condition confirmed the identical
behavioral pattern of stress-strain and stress-acoustic emission curves at all levels of confining
pressure, where the emissivity of acoustic signals varies with different shape and gradation of the
particles (Koerner et al., 1976). The acoustic emission events of greater magnitude are produced
in granular soil with large angular particles and with high interparticle contact stress (Koerner et
al., 1976). Pioro and Oshkin (2011) confirms that both longitudinal and shear velocities are
higher in sand particles than that in clay. This may be because of the high attenuation of the
porous particles.

CONCLUSION

Since AE is becoming a widely accepted monitoring technology, it also has huge potential in
the field of geotechnical engineering. Correlating the various physical parameters of the soil

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 241

deformation with the acoustic emission parameters, it is possible to design a system that could
provide early information regarding the undergoing situation of soil deformation. Various
research studies have shown the use of AE in case of slope stability but the studies also showed
that it has potential to be used in other systems of geotechnical engineering like foundations of
the buildings. However, developing widely applicable AE interpretation procedures that decision
makers may use to deliver information on the status of a variety of geotechnical assets is still a
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

challenging job. In order to establish a widely accepted early design system, further research
investigations are needed in this sector.

ACKNOWLEDGEMENTS

This material is based upon work supported by the Broad Agency Announcement Program
and the Cold Regions Research and Engineering Laboratory (ERDC-CRREL) under Contract
No. W913E522C0001. Any opinions, findings and conclusions or recommendations expressed in
this material are those of the author(s) and do not necessarily reflect the views of the Broad
Agency Announcement Program and ERDC-CRREL.

REFERENCES

Berg, N., Smith, A., Russell, S., Dixon, N., Proudfoot, D., and Take, W. A. 2018. “Correlation of
acoustic emissions with patterns of movement in an extremely slow-moving landslide at
Peace River, Alberta, Canada.” Canadian Geotechnical Journal, 55: 1475-88.
Caicedo, B., and Patino-Ramirez, F. 2019. Acoustic emission instrumentation method for slope
monitoring. In E3S Web of Conferences, 18010. EDP Sciences.
Cheon, D.-S., Jung, Y.-B., Park, E.-S., Song, W.-K., and Jang, H.-I. 2011. “Evaluation of
damage level for rock slopes using acoustic emission technique with waveguides.”
Engineering Geology, 121: 75-88.
Codeglia, D., Dixon, N., Fowmes, G. J., and Marcato, G. 2017. “Analysis of acoustic emission
patterns for monitoring of rock slope deformation mechanisms.” Engineering Geology, 219:
21-31.
Deng, L., Yuan, H., Chen, J., Sun, Z., Fu, M., Zhou, Y., Yan, S., Zhang, Z., and Chen, T. 2019.
“Experimental investigation on progressive deformation of soil slope using acoustic emission
monitoring.” Engineering Geology, 261: 105295.
Dixon, N., Hill, R., and Kavanagh, J. 2003. “Acoustic emission monitoring of slope instability:
development of an active waveguide system.” Proceedings of the institution of civil
engineers-geotechnical engineering, 156: 83-95.
Dixon, N., Smith, A., Spriggs, M., Ridley, A., Meldrum, P., and Haslam, E. 2015. “Stability
monitoring of a rail slope using acoustic emission.” Proceedings of the Institution of Civil
Engineers-Geotechnical Engineering, 168: 373-84.
Dixon, N., and Spriggs, M. 2007. “Quantification of slope displacement rates using acoustic
emission monitoring.” Canadian Geotechnical Journal, 44: 966-76.
Dixon, N., Spriggs, M. P., Smith, A., Meldrum, P., and Haslam, E. 2015. “Quantification of
reactivated landslide behaviour using acoustic emission monitoring.” Landslides, 12: 549-60.
Koerner, R. M., Curran, J. W., Mccabe, W. M., and Lord, A. E., Jr. 1976. “Acoustic emission
behavior of granular soils.” Journal of the Geotechnical Engineering Division, 102: 761-73.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 242

Koerner, R. M., Mccabe, W. M., and Lord, A. E., Jr. 1978. “Acoustic emission monitoring of
soil stability.” Journal of the Geotechnical Engineering Division, 104: 571-82.
Lin, W., Liu, A., and Mao, W. 2019. “Use of acoustic emission to evaluate the micro-mechanical
behavior of sands in single particle compression tests.” Ultrasonics, 99: 105962.
Mao, W., Aoyama, S., Goto, S., and Towhata, I. 2016. “Behaviour and frequency characteristics
of acoustic emissions from sandy ground under model pile penetration.” Near Surface
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Geophysics, 14: 515-25.


Mao, W., Yang, Y., Lin, W., Aoyama, S., and Towhata, I. 2018. “High Frequency Acoustic
Emissions Observed during Model Pile Penetration in Sand and Implications for Particle
Breakage Behavior.” International Journal of Geomechanics.
Michlmayr, G., Cohen, D., and Or, D. 2013. “Shear‐induced force fluctuations and acoustic
emissions in granular material.” Journal of Geophysical Research: Solid Earth, 118: 6086-
98.
Pioro, E. V., and Oshkin, A. N. 2011. “Relationships between the acoustic characteristics and
physical and deformation properties of clay soils.” Moscow University Geology Bulletin, 66:
449-52.
Smith, A., and Dixon, N. 2020. “Listening for deterioration and failure: towards smart
geotechnical infrastructure.” Proceedings of the Institution of Civil Engineers-Smart
Infrastructure and Construction, 171: 131-43.
Smith, A., Dixon, N., and Fowmes, G. 2017. “Monitoring buried pipe deformation using acoustic
emission: quantification of attenuation.” International Journal of Geotechnical Engineering,
11: 418-30.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 243

An Artificial Neural Network Model for Predicting Microbial-Induced Alteration of Rock


Strength

Oladoyin Kolawole, Ph.D.1; Rayan H. Assaad, Ph.D.2; Mary C. Ngoma3;


and Ogochukwu Ozotta, Ph.D.4
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Assistant Professor, Dept. of Civil and Environmental Engineering, New Jersey Institute of
Technology, Newark, NJ (corresponding author). Email: oladoyin.kolawole@njit.edu
2
Assistant Professor, Dept. of Civil and Environmental Engineering, New Jersey Institute of
Technology, Newark, NJ. Email: rayan.hassane.assaad@njit.edu
3
Graduate Student, Dept. of Civil and Environmental Engineering, New Jersey Institute of
Technology, Newark, NJ. Email: mcn@njit.edu
4
Associate, Dept. of Petroleum Engineering, Univ. of North Dakota, Grand Forks, ND.
Email: ogochukwu.ozotta@und.edu

ABSTRACT

In bio-mediated geotechnical techniques, the estimation of microbially altered geomechanical


properties has mostly been conducted at core-scale due to the complexity of in situ field-scale
measurement of biogeomechanical properties. However, the successful in situ field application
of this emerging field of geomechanics relies on proper upscaling of this process from core-scale
to field-scale (reservoir-scale). Here, we developed and applied a machine learning (ML)
algorithm (artificial neural network, ANN) to model and predict the reservoir-scale
biogeomechanical-altered properties of shale and carbonate rocks. We first obtained
experimental data of the core- and bulk-scale mechanical properties (uniaxial compression
strength, UCS) of the core samples impacted by a microbial strain. These core-scale data were
then subsequently used as input variables to predict the field-scale biogeomechanical altered
properties. The results show a high degree of correlation between the ML-predicted field-scale
biogeomechanical properties and the laboratory-obtained bulk-scale biogeomechanical properties
in shales (11.2% mean absolute percentage error) and carbonates (13.5% mean absolute
percentage error). In addition, the result shows that the degree of correlation in rock mechanical
properties and new mineral precipitations may be higher with increasing pore spaces in the tested
rock types. This study provides a first leap from the laboratory and core-scale investigations
toward field-scale geotechnical and geo-environmental applications of biocementation and
biomineralization by predicting in situ biogeomechanical alterations.

INTRODUCTION

Biogeomechanics, an emerging field of biogeotechnics, investigates the geomechanical


responses of microbial-rock interactions (Kolawole et al. 2021a, 2021b, 2022a, 2022b).
Biomineralization and bio-inspired approaches have been adopted over the years to tackle
problems of geotechnical interest.
Biocementation techniques have been adopted for slope and soil stabilization and
improvement (Wang et al. 2019; Oualha et al. 2020; Terzis et al. 2020), hydraulic control
(Phillips et al. 2013), and erosion control (Jiang et al. 2017). Similarly, in rocks and rock-like
materials, biofilm application has been explored for the improvement of wellbore integrity

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 244

(Phillips et al. 2018; Kirkland et al. 2020), geological CO2 sequestration (Kolawole et al. 2021b,
2022a; Landa-Marbán et al. 2021), building improvement (Dhami et al. 2013; Gao et al. 2019),
and hydrocarbon recovery improvement (Kolawole et al. 2021c, 2022b). Previous studies have
experimentally investigated and observed the micro-scale (Figure 1; Figure 2) impact of
biological processes on geomaterials (Terzis et al. 2020; Kolawole et al. 2021a).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Micro-scale observations of biocementation in shale rock (Kolawole et al. 2021a).

Figure 2. Micro-scale observations of biocementation in clay soil (Terzis et al. 2020).

To break out of the laboratory investigation limitations, it is critical to upscale laboratory


results to deliver large-scale bio-cementation of geomaterials in near-surface and subsurface
environments. Also, the investigations of microbially-altered geomechanical properties have
mostly been conducted at the core-scale (micro- and meso-scale) due to the complexity of in-situ
field-scale measurement of biogeomechanical properties. Therefore, the successful in-situ field
application of this emerging field of geomechanics relies on proper upscaling of this process
from core-scale to field-scale (reservoir-scale) geomechanical characterization.
The goal of this study is to investigate the feasibility of a machine learning (ML) algorithm to
model and predict the reservoir-scale biogeomechanical-altered properties of shale and carbonate
rocks.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 245

MATERIALS AND METHODS

Laboratory data from biogeomechanical investigations on shale and carbonate rocks were
used for this study, as experimental data. We first obtained laboratory data of the core- and bulk-
scale geomechanical properties (uniaxial compression strength, UCS) of the rock samples in the
presence or absence of the microbial agent. The procedure for the microbial treatment of the rock
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

samples is well detailed in Kolawole et al. (2021a, 2022a). Further, an Artificial Neural Network
(ANN) machine learning algorithm was developed. ANNs are considered a family of non-linear
computational models or algorithms that were inspired by the human brain and have been
implemented in an increasing number of applications and fields (Marini, 2009). In simple words,
an ANN model is a collection of computational units (called cells or neurons) that are interlinked
by a system of weighted connections (also known as synaptic connections) (Kanevsky, 1996).
ANNs were used in this study because they are considered a very powerful algorithm that can
represent non-linear trends within the data and model highly complex relationships, which is the
case of rock mechanical properties.
The collected core-scale data were subsequently used as input variables to predict the field-
scale biogeomechanical altered properties. Other data used in developing the ANN algorithm
include rock type, microbe specimen type and culture volume, growth nutrient, treatment
temperature and pressure conditions, microbial media injection rate and pressure, the volume of
microbial fluid injected, and the microbial treatment time.
The collected data was divided into 70% as training data and 30% as testing data. The testing
data was held out for the final evaluation of the selected and developed ANN model on unseen
data to ensure its robustness. k-fold cross-validation was performed on the training data by
splitting it into k folds (or k parts) where the model is trained on k-1 folds and its accuracy is
reported or validated on the remaining cross-validation fold. This is repeated k times to ensure
that the cross-validation set was run on the entire training data. More specifically, 5-fold cross-
validation was used in this study, which is commonly used when developing machine learning
models because it ensures the generalization and robustness of the developed machine learning
models and avoids overfitting (Hong et al. 2018; Chemchem et al. 2019).
The inputs to the developed ANN include: diameter, pre-treatment weight, pre-treatment
density, treatment period, treatment temperature, treatment pressure, microbial injection rate,
volume of microbial sample added, volume of growth/nutrient used, volume of additional
materials included in microbial solution, total volume of microbial solution used for treatment,
length, pre-UCS, pre-Poisson's ratio, and microbial media. Before training and developing the
ANN model, the min-max scaling method was used to bring all input variables or features to the
[0, 1] range because the collected input variables have varying magnitudes, units, and ranges.
This creates magnitude issues and thus there is a need to bring all features to the same level of
magnitudes (Assaad and El-adaway 2020; Elsayegh et al. 2020). The min-max scaling method is
shown in Equation 1.

𝑥−min(𝑥)
𝑥𝑠𝑐𝑎𝑙𝑒𝑑 = (1)
max(𝑥)−min(𝑥)

Where 𝑥𝑠𝑐𝑎𝑙𝑒𝑑 is the scaled version of the 𝑥 variable/feature, min(𝑥) is the minimum value of
the variable/feature, and max(𝑥) is the maximum value of the variable/feature.
In ML, the training process involves determining the optimal hyperparameters that the
learning algorithm will use to best map, and model the relationship between, the input features

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 246

(independent variables) to the labels or targets (dependent variable). Thus, a hyperparameter is a


parameter whose value is used to control the learning process. In relation to that, different
combinations of hyper-parameters were considered when developing the ANN model as shown
in Table 1.

Table 1. Considered hyper-parameters for the ANNs machine learning algorithm


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Algorithm Considered hyper-parameters


ANN • Number of hidden layers 1, 2, 3, 4, 5, and 6
• Number of neurons 1, 2, 3, 4, 5
• Learning rate 1, 0.9, 0.8, 0.7, 0.6, 0.5, 0.4, 0.3, 0.2, 0.1, 0.01, and 0.001
• Size of mini-batches: 16, 36, 64, 128, and 256
• L2 regularization (i.e., alpha): 0.0001 and 0.05

RESULTS AND DISCUSSION

The ANN algorithm was used to predict the post-treatment UCS of shale rocks and carbonate
rocks. Different hyper-parameters were used as details in the “Materials and Methods” section.
The graphs of predicted values vs actual values for the shale rocks and carbonate rocks are
shown in Figure 3 and Figure 4, respectively, with the 45-degree line represented in dotted-red. It
is to be noted that data points close to the red-dotted line represent accurate predictions (i.e., the
predicted value of UCS is close to the actual value of UCS).
To better quantify the error metrics associated with the graphs shown in Figures 3 and 4, the
performance metrics on the testing set for the ANN algorithm for shale rocks and carbonate
rocks are shown in Tables 2 and 3.

Table 2. ANN Prediction of Biogeomechanical Alteration in Shale Rocks

Attribute Value
Rock type Shale
Best hyper-parameters ‘alpha’ is 0.05, ‘batch size’ is 128, ‘hidden
layer sizes’ is (3, 3, 3), ‘learning rate’ is 0.5,
and 'solver’ is 'adam’.
Mean absolute error (MAE) 10.11 MPa
Mean error (ME) 3.207 MPa
Root mean square error (RMSE) 12.49 MPa
Mean absolute percentage error (MAPE) 11.19%

In shale rocks (Table 2; Figure 3), the ANN model shows that the UCS mean absolute error is
10.11 MPa, and the UCS mean absolute percentage error is 11.19%. The best hyper-parameters
are 'alpha': 0.05, 'batch_size': 128, 'hidden_layer_sizes': (3, 3, 3), 'learning_rate_init': 0.5, and
'solver': 'adam'. The UCS mean error in shales is 3.21 MPa. This suggests that ANNs can

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 247

accurately predict rock strength modification in shale rocks due to the biological process with a
low error and high correlation.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Analysis of ANN model prediction and actual biogeomechanical alterations in


shale rocks.

Table 3. ANN Prediction of Biogeomechanical Alteration in Carbonate Rocks

Attribute Value
Rock type Carbonate
Best hyper-parameters 'alpha’ is 0.05, 'batch size’ is 36,
‘hidden layer sizes’ is (4, 4), 'learning
rate’ is 0.4, and 'solver’ is 'adam’.
Mean absolute error (MAE) 11.26 MPa
Mean error (ME) 0.287 MPa
Root mean square error (RMSE) 14.99 MPa
Mean absolute percentage error (MAPE) 13.46%

In carbonate rocks (Table 3; Figure 4), the ANN model prediction shows that the UCS mean
absolute error is 11.26 MPa, and the UCS mean absolute percentage error is 13.46%. The best
hyper-parameters are 'alpha': 0.05, 'batch_size': 36, 'hidden_layer_sizes': (4, 4),
'learning_rate_init': 0.4, and 'solver': 'adam'. The UCS mean error in carbonates is 0.29 MPa. This

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 248

suggests that in carbonates, the ANN model can predict rock strength modification as a result of
the biological process with a low error and high correlation.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Analysis of ANN model prediction and actual biogeomechanical alterations in


carbonate rocks.

The obtained results in Figures 3 and 4 reflected that the predicted values are more dispersed
for the shale rocks compared to those of the carbonate rocks, which could be attributed to the
inherent difference in the distributions of the rock strengths and properties (UCS in particular)
between shales and carbonates as shown in Figures 5 and 6, respectively. In relation to that,
Figure 6 shows that the carbonate rocks have a multi-modal model distribution (more
specifically, a bimodal distribution) whereas Figure 5 shows that the shale rocks have a unimodal
distribution.
These differences in the properties of the distributions of the two types of rocks reflect the
complexity and inherent challenge of being able to develop accurate models that can precisely
predict or estimate the strengths of the two types of rock post-microbial treatment. Nevertheless,
and despite the aforementioned differences and inherent modeling complexity, the developed
ANN models provided comparable prediction performances between the two types of rocks (i.e.,
mean absolute percentage error of 11.19% for shale rocks vs mean absolute percentage error of
13.46% for shale rocks).
This reflects that no matter how complex and non-linear the potential relationship is between
the input and output variables, as long as the samples and structures of the ANN model are
properly and reasonably designed and optimized through hyper-parameter tuning, the ANN
algorithm can provide excellent flexibility and adaptability to approximate non-linear models
with high accuracy (or low error measures) as Tables 2 and 3 have shown. In fact, the flexibility

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 249

and adaptability of the developed ANN models are reflected by the fact the ANN model for the
shale rocks has different hyper-parameters than the ANN model for the carbonate rocks (i.e., 3
hidden layers for shales vs 2 hidden layers for carbonates, 3 neurons in each of the hidden layers
for shales vs 4 neurons in each of the hidden layers for carbonates, learning rate of 0.5 for shales
vs learning rate of 0.4 for carbonates, and batch size of 128 for shales vs batch size of 36 for
carbonates).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5. Distribution of the UCS for shale rocks (values from experimental observations).

Figure 6. Distribution of the UCS for carbonate rocks (values from experimental
observations).

In summary, after incorporating the pre- and post-microbial treatment data and conditions in
the ANN algorithm, the results suggest the great potential of using ANNs for biogeomechanical
prediction in shale rocks (Table 2) and carbonate rocks (Table 3). Thus, the biogeomechanical
analyses of the ANN model results (predicted) and the experimentally obtained bulk-scale
(actual) properties in shales (Figure 3) and carbonates (Figure 4) indicate a high degree of
correlation between the predicted field-scale UCS and the laboratory-obtained bulk-scale UCS in
shale rocks (mean absolute percentage error = 11.2%) and carbonate rocks (mean absolute
percentage error = 13.5%).
The analyses of our results indicate a higher mean error in shales relative to carbonates. Since
the microbes are expected to grow through the pore spaces (containing the injected growth
nutrients) in rocks and induce mineral deposits and cementation during the microbial treatment

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 250

process. Thus, we suspect the higher mean error in shales (3.2 MPa) might be due to its
characteristic ultra-low porosity and permeability in comparison with carbonate rocks (0.29
MPa) which are usually more porous and permeable.

CONCLUSIONS
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

This work developed and applied an Artificial Neural Network (ANN) model to predict the
reservoir-scale biogeomechanical-altered properties (UCS) of shale and carbonate rocks. The
results show a high degree of correlation between the ML-predicted field-scale
biogeomechanical properties and the laboratory-obtained bulk-scale biogeomechanical properties
in shales (11.2% mean absolute percentage error) and carbonates (13.5% mean absolute
percentage error). In addition, the results show that the degree of correlation in geomechanical
properties and new mineral precipitations may be higher with increasing pore spaces in the
laboratory-tested rock types.
Overall, the novel work presented herein attempted a first leap from the laboratory and core-
scale investigations toward field-scale geotechnical and geo-environmental applications of
biocementation and biomineralization by predicting in-situ biogeomechanical alterations.
Although this current work is only limited to ANN and rock strength, however, the next
phase of this project is to develop of more robust ML approach which incorporates other ML
algorithms and more laboratory data and more mechanical properties to validate and increase the
accuracy of the ML models.

REFERENCES

Assaad, R., and El-adaway, I. H. (2020). “Evaluation and prediction of the hazard potential level
of dam infrastructures using computational artificial intelligence algorithms.” Journal of
Management in Engineering, 36(5), 04020051.
Chemchem, A., Alin, F., and Krajecki, M. (2019). “Combining SMOTE sampling and machine
learning for forecasting wheat yields in France.” In Proc., 2019 IEEE 2nd Int. Conf. on
Artificial Intelligence and Knowledge Engineering (AIKE), 9–14. New York: IEEE.
10.1109/AIKE.2019.00010.
Dhami, N. K., Reddy, M. S., and Mukherjee, A. (2013). “Biomineralization of calcium
carbonates and their engineered applications: a review.” Frontiers in Microbiology.
Elsayegh, A., El-Adaway, I. H., Assaad, R., Ali, G., Abotaleb, I., Smith, C., Bootwala, M., and
Eteifa, S. (2020). Contractual guidelines for management of infrastructure transportation
projects. Journal of Legal Affairs and Dispute Resolution in Engineering and Construction,
12(3), p.04520023.
Gao, R., Luo, Y., and Deng, H. (2019). “Experimental study on repair of fractured rock mass by
microbial induction technology.” R Soc Open Sci, 6:11.
Hong, H., Zhu, J., Chen, M., Gong, P., Zhang, C., and Tong, W. (2018). “Quantitative structure–
activity relationship models for predicting risk of drug-induced liver injury in humans.” In
Drug-induced liver toxicity, 77–100. New York: Humana Press. 10.1007/978-1-4939-7677-
5_5.
Jiang, N.-J., Soga, K., and Kuo, M. (2017). “Microbially induced carbonate precipitation for
seepage- induced internal erosion control in sand-clay mixtures.” J. Geotech. Geoenviron.
Eng. 143 (3): 04016100.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 251

Kanevsky, M., Arutyunyan, R., Bolshov, L., Demyanov, V., and Maignan, M. (1996). “Artificial
neural networks and spatial estimation of Chernobyl fallout.” Geoinformatics, 7(1-2), 5-11.
Kirkland, C. M., et al. (2020). “Direct Injection of Biomineralizing Agents to Restore Injectivity
and Wellbore Integrity.” SPE Production & Operations, 36 (01): 216-223.
Kolawole, O., Millikan, C., Kumar, M., Ispas, I., Schwartz, B., Weber, J., Badurina, L., and
Šegvić, B. (2022a). “Impact of microbial-rock-CO2 interactions on containment and storage
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

security of supercritical CO2 in carbonates.” International Journal of Greenhouse Gas


Control, 120(4), 103755.
Kolawole, O., Ispas, I., Kumar, M., Weber, J., and Zhao, B. (2021a). “Time-Lapse
Biogeomechanical Modified Properties of Ultra-Low Permeability Reservoirs.” Rock
Mechanics and Rock Engineering, 54, 2615–2641.
Kolawole, O., Ispas, I., Kumar, M., and Huffman, K. (2021c). “Biogeomechanical Alteration of
Near- Wellbore Properties: Implications for Hydrocarbon Recovery.” Journal of Natural Gas
Science and Engineering, 94, 104055.
Kolawole, O., Ispas, I., Kumar, M., Weber, J., Zhao, B., and Zanoni, G. (2021b). “How Can
Biogeomechanical Alterations in Shales Impact Caprock Integrity and CO2 Storage?.” Fuel,
291, 120149.
Kolawole, O., Millikan, C., Kumar, M., Ispas, I., and Weber, J. (2022b). “Microbial Induced
Mechano-Petrophysical Modified Properties to Improve Hydrocarbon Recovery in Carbonate
Reservoirs.” Geomechanics for Energy and the Environment, 100399.
Landa-Marbán, D., et al. (2021). “Practical approaches to study microbially induced calcite
precipitation at the field scale.” International Journal of Greenhouse Gas Control, 106.
Marini, F. (2009). “Artificial neural networks in foodstuff analyses: Trends and perspectives A
review.” Analytica Chimica Acta, 635(2), 121-131.
Oualha, M., et al. (2020). “Microbially induced calcite precipitation in calcareous soils by
endogenous Bacillus cereus, at high pH and harsh weather.” Journal of Environmental
Management, 257.
Phillips, A. J., et al. (2018). “Enhancing wellbore cement integrity with microbially induced
calcite precipitation (MICP): A field scale demonstration.” Journal of Petroleum Science and
Engineering, 171, 1141-1148.
Phillips, A. J., Lauchnor, E., Eldring, J., Esposito, R., Mitchell, A. C., Gerlach, R., and Spangler,
L. H. (2013). “Potential CO2 leakage reduction through biofilm-induced calcium carbonate
precipitation.” Environ. Sci. Technol., 47 (1): 142–149.
Terzis, D., Laloui, L., Dornberger, S., and Harran, R. (2020). “A Full-Scale Application of Slope
Stabilization via Calcite Bio-Mineralization Followed by Long-Term GIS Surveillance.”
Proc., Geo-Congress, Minneapolis, MN, USA.
Wang, Y., Soga, K., DeJong, J. T., and Kabla, A. J. (2019). “Microscale Visualization of
Microbial- Induced Calcium Carbonate Precipitation Processes.” J. Geotech. Geoenviron.
Eng., 145(9): 04019045.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 252

Investigating Influence of Freeze-Thaw Cycles on Sandstone Containing Pre-Existing


Joints through Discrete Element Modeling

Chenchen Huang1; Cheng Zhu, Ph.D., P.E.2; Yifei Ma, Ph.D., P.E.3;
and Shaini Aluthgun Hewage, Ph.D.4
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Graduate Research Assistant, Dept. of Civil and Environmental Engineering, Rowan Univ.,
Glassboro, NJ. Email: huangc57@students.rowan.edu
2
Assistant Professor, Dept. of Civil and Environmental Engineering, Rowan Univ., Glassboro,
NJ. Email: zhuc@rowan.edu
3
Assistant Professor, Dept. of Civil and Architectural Engineering, Lawrence Technological
Univ., Southfield, MI. Email: yma@ltu.edu
4
Postdoctoral Researcher, Dept. of Civil and Environmental Engineering, Rowan Univ.,
Glassboro, NJ. Email: aluthgunhewage@rowan.edu

ABSTRACT

In this study, the discrete element method (DEM) is adopted to evaluate the damage and
cracking behaviors of sandstone containing two unparallel fissures under freeze-thaw cycles. The
rock and water compositions are represented by particles of different sizes and properties, with
smaller water particles filled into rock pores. A modified elastoplastic parallel bonded model is
developed to describe the accumulation of residual strain caused by the ice-water phase change.
The model is calibrated by comparing numerical results [uniaxial compressive strength (UCS),
Young’s modulus, and failure pattern] with published experimental data. Compared with
published experimental results, similar crack coalescence patterns and similar deteriorating
effects of freeze-thaw cycles (FTCs) are observed in the numerical results. This study enhances
our understanding of the influence of FTC on the cracking behaviors of sandstone containing
fissures and may contribute to the future design and assessment of climate-resilient
infrastructure.

INTRODUCTION

Many discontinuities (e.g., joints, faults, and fissures) in rock mass greatly influence the
physical and mechanical properties of rock (Wang et al. 2020). Especially in regions undergoing
gradual warming or freeze-thaw cycles, the existence of discontinuities exacerbates the
deteriorative influence of such temperature variations on rocks, as a result of the water phase
transition within these discontinuities and the induced volume change and local stress
concentration (Lu et al. 2019). Therefore, understanding the damage mechanism of rock
containing cracks under freeze-thaw cycles is significant.
To explore the mechanical behaviors and cracking mechanisms of flawed rocks, some studies
have used experimental methods to evaluate the mechanical properties and failure patterns of
rocks or rock-like material containing pre-existing fissures (Huang et al. 2019). Previous studies
mainly focused on the cracking behaviors of rock or rock-like materials containing pre-existing
joints at room temperature. Some studies (Lu et al. 2019; Niu et al. 2019; Wang et al. 2020; Zhou
et al. 2019) investigated the mechanical and cracking behaviors of flawed rock subjected to
FTCs. For example, Zhou et al. (2019) investigated the influence of FTCs on the cracking
behaviors and mechanical parameters (e.g., UCS, initiation stress, elastic modulus, and peak

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 253

axial strain) of sandstone specimens with two unparallel fissures under uniaxial loading. Niu et
al. (2019) analyzed the fracturing characteristics of unparallel fissure-contained sandstone
specimens subject to different FTCs under uniaxial compression. The cracking processes of
specimens were investigated based on the high-speed digital video camera system.
Most existing studies are based on laboratory tests, whereas the underlying mechanism of
cracking behaviors (crack initiation, propagation, and coalescence) responsible for the FTCs-
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

induced degradation remains unclear. Existing numerical studies based on Discrete Element
Method (DEM) (Zhu et al. 2021) on FTCs relied on linear models and did not consider the pre-
existing joints in specimens. In this study, we developed an elastoplastic parallel bond model in
this study to capture the micro-damage in jointed sandstone specimens during FTCs. Combined
with the smooth-joint model, the influence of FTCs on the mechanical properties of sandstone
with two unparallel joints was analyzed in this study. The study contributes to the understanding
of the influences of FTCs on sandstone with joints in a changing climate and might help with the
assessment of climate-resilient infrastructure in the future.

METHODOLOGY

Numerical Model of Sandstone

The DEM software, Particle Flow Code in Two Dimension (PFC2D), was used in this
research to explore the influence of FTCs on the sandstone specimens containing two unparallel
joints. Discrete elements interact with neighboring elements using classical equations of motion.
The trial-and-error method was used to ascertain the micro parameters used in DEM models
(Huang et al. 2019).
Referred to the experimental data from Niu et al. (2019) as shown in Figure 1(a), a numerical
model (height = 160 mm, width = 80 mm) was constructed, with sandstone particles ranging
from 0.7 to 1.05 mm and water particles (one water particle represents a cluster of water
molecules) from 0.5 to 0.75 mm as illustrated in Figure 1(b). This numerical model contains
4594 rock particles (grey circles) and 482 water particles (blue circles). Because the basic
discrete elements in PFC2D are circular, it is difficult to directly consider the porosity of an actual
rock sample accurately . The porosity is assumed as the ratio between all water particles’ volume
and all rock particles’ volume in this study. To make water particles better interact with
surrounding rock particles, a filling method based on FISH programming was developed to add
more water particles in the pores of the numerical sample. Firstly, a random point inside the
original rock sample was found by the FISH built-in function. Then, we compute the distance
between the random point and each existing particle. If the distances between the random point
and each surrounding particle are greater than the radius of the corresponding particle, it is
confirmed that the random point was located inside the pore space. After that, a water particle
will be generated at this random point (Huang et al. 2022). This filling method was performed
more than 200,000 times to add enough water particles until the target porosity was achieved.
To produce the pre-existing joints in the DEM model, some previous studies deleted the
particles corresponding to the joint positions to simulate the non-closed joints . However, at least
one layer of particles is deleted in this method which would result in roughness difference and
thickness difference of pre-existing joints with experimental conditions . The smooth-joint model
can be used to overcome the roughness problems . As Figure 1(c) shows, after pre-existing joints
(green lines) are added, surrounding contacts were replaced by smooth-joint contact model
(black lines). The thickness of pre-exsiting joints is 0.016 mm.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 254
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Geometry of pre-existing joints and sample size used in this study: (a) sandstone
sample used in experiments from Niu et al. (2019), in which 2a is the joint length, 2b is the
ligament length, α is the joint angle, and β is the ligament angle; (b) DEM model built in
this study, in which the grey circles are rock particles, blue circles are water particles,
black lines are pre-existing joints; (c) contact models used in this study (black lines denote
the smooth-joint model, red lines denote the elastoplastic parallel bond model, and light
blue lines denote the linear parallel bond model).

The water-water (W-W) and water-rock (W-R) contacts were assigned with the linear
parallel bond model (white lines in Figure 1(c)), which accounts for the phase transition of water
into ice and the resulting rock-like rigid behavior (Huang et al. 2019). A developed elastoplastic
parallel bond model (red lines in Figure 1(c)) used for rock-rock (R-R) contacts will be
introduced in the following section.

Elastoplastic Parallel Bond Model

Considering the damage caused by FTCs in the linear contact model would recover in the
thawing stage before the bond break, we proposed an elastoplastic parallel bond contact model to
capture the damage of sandstone samples caused by FTCs based on C++ programming. The
normal force-displacement curve of the elastoplastic parallel bond model can be divided into
three stages as shown in Figure 2: elastic stage, plastic softening stage, and plastic flow stage.
The normal force in three stages can be calculated as Eq. (1):

 kne Ag g  gss



Fn = −knp Ag gss  g  g fs (1)
 0 g fs  g

where kne is normal elastic stiffness, knp is normal plastic stiffness, gss is softening
displacement, g fs is flow displacement, g is the relative normal displacement increment, and
A is the area of the bond cross-section.
The Mohr-Coulomb criteria was used to determine the shear-strength limit as shown in Eq.
(2):

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 255

c = c −  tan  (2)

where c is shear strength, c is cohesion,  is normal strength, and  is friction angle. The
bond breaks in shear if the shear stress surpasses pre-defined shear strength.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Normal force-displacement curve of elastoplastic parallel bond model.

Thermal Contact and Thermal Properties

The built-in thermal contacts in PFC2D were used to achieve heat conduction in the sample and
calculate the thermal expansion of water phase transition. The User’s manual introduces the
mechanism of thermal calculation. Each particle was regarded as a reservoir for storing heat in
thermal contacts. Heat conduction occurs when there is a temperature difference between
neighboring particles. The heat flux can be calculated by Fourier’s heat conduction law as shown
in Eq. (3):

T
q = −k (3)
x

where q is the heat flux, k is the thermal conductivity, and T x is the temperature gradient
between two neighboring particles.
Eq. (4) shows the thermal expansion of particles ( R ) caused by heat conduction.

R =  RT (4)

where  is the thermal expansion coefficient of the material, and T is the temperature
increment.
Assuming the expansion of bond material is isotropic in the thermal calculation, Eq. (5) can
be used to calculate the change of normal force caused by the expansion:

Fn = −kn AUn = −kn A( L T ) (5)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 256

where Fn is the normal force change of elastoplastic parallel bond contact, k is the normal
stiffness of bond, A is the area of bond cross-section,  is the expansion coefficient of the bond
that is equal to the average value of expansion coefficient of two particles, and L is the bond
length.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Boundary and Initial Conditions

Considering the fractures of rocks terminated development at about -5 ℃ during the freeze-
thaw experiments , the temperature interval was set from 0 ℃ to -10 ℃ in order to improve the
calculation efficiency in this study.
Because the water is frozen as ice whose volume increases by about 9% when the
temperature is below 0 ℃ , the thermal expansion coefficient was set as 0.0044 m/℃ to account
for the volume change of water phase transition. This expansion would enlarge the pore spaces
and cause an increase of rock volume from the macro perspective, which allows more water to
enter the pores after ice thawing. Considering the extremely small volume of pores, adding more
small water particles in these pores would decrease the calculation efficiency. Therefore, the
thermal expansion coefficient of ice was set as 0.0037 m/℃ to simulate more water particles
added to the sample (Khanlari et al. 2014). The volume change of rock-solid was negligible
compared with the volume change of water produced by phase transition.
As shown in Figure 3(a), three groups of particles (boundary rock particles, center particle
and other particles) were generated for thermal calculation. For example, The temperature of all
particles was set as 0 ℃ before FTCs. During the freezing stage, the temperature of boundary
rock particles was fixed at -10 ℃ to simulate the low ambient temperature condition. The heat
conduction as shown in Figure 3(b) occurred until the temperature of the center particle
decreases to -10 ℃. After that, the temperature of boundary particles was set as 0 ℃ for thawing.

Figure 3. Thermal calculation model: (a) group particles to apply boundary conditions; (b)
Temperature (℃) distribution freezing stage.

MODEL CALIBRATION AND VALIDATION

By carrying out the uniaxial compressive test on the sandstone samples with pre-existing
joints subjected to the different number of freeze-thaw cycles (i.e.,0, 15), the micro-parameters

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 257

(i.e., modulus, stiffness ration, soften strain, flow strain, and peak strain) of the elastoplastic
parallel bond model were calibrated. To be consistent with the dry condition in experiments, all
water particles were deleted before carrying out the uniaxial compressive test. A constant
displacement rate of 0.005 m/s was applied to the samples in a displacement-controlled way.
Huang and Zhu (2018) have proved that this rate is low enough to ensure a quasi-static numerical
loading process.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

By comparing the numerical results with experimental results from Niu et al. (2019), the
micro-parameters listed in Table 1 were calibrated. The additional numerical uniaxial
compressive tests on the sandstone samples treated by 30, 45, and 60 FTCs were performed to
validate the calibrated micro-parameters.

(a) (b)

Figure 4. Comparison of the mechanical properties obtained on the experiments from Niu
et al. (2019) and numerical simulation: (a) UCS; (b) Young’s modulus.

Figure 4 shows the comparison of UCS and Young’s modulus obtained from the experiments
and numerical simulation. With the FTCs increasing, the UCS and Young’s modulus of
sandstone sample with pre-existing joints decrease. The numerical results show good agreement
with the published experimental data (Niu et al. 2019), which validates that the developed
elastoplastic parallel bond model can be used to capture the deteriorative influence of FTCs on
the mechanical properties of sandstone sample.

NUMERICAL SIMULATION RESULTS

The Influence of FTCs on Mechanical Properties of Sandstone Samples

Axial stress-strain curves of sandstone samples containing two unparallel joints subjected to
different FTCs under compression are shown in Figure 5. It is clear that the FTCs have a great
effect on the strength and deformation behavior of sandstone samples. The uniaxial compressive
strength and Young’s modulus decrease with the increasing number of FTCs. The nonlinear
deformation of the sandstone samples becomes larger as the number of FTCs increases, which
can be explained by the fact that the temperature change and water phase transition enlarge the
pores and voids in the sandstone samples containing two unparallel joints (Niu et al. 2019). The
deteriorative influences of FTCs on the sandstone samples were explored by analyzing the
microstructure of sandstone samples subjected to FTCs in the following section.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 258

Table 1. Micro-parameters for numerical sandstone sample with (the parameters of W-R
bond and W-W bond are adopted from Zhu et al. (2021)).

Parameter of parallel-bond model R-R bond W-R bond W-W bond


Elastic modulus (GPa) 5 5 5
Plastic modulus (GPa) 4 - -
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Stiffness ratio 1.49 1.49 1.5


Tensile strength (MPa) - 100 100
Cohesion (MPa) 80 100 100
Soften strain (10-3) 0.35 - -
Flow strain (10-3) 0.5 - -
Peak strain (10-3) 2.8 - -
Friction (°) 0 0 0
Parameter of particle Rock particle Water particle
Minimum particles radius (mm) 0.70 0.5
Maximum particles radius (mm) 1.05 0.75
Friction coefficient 0.4 0.01
Stiffness ratio 1.5 1.5
Density (kg/m3) 2460 980
Parameter of smooth-joint model
Normal/Shear strength (MPa) 0
Normal/Shear stiffness (N/m) 2×109
Friction coefficient 0.5

Figure 5. Stress-strain curves of sandstone samples containing two unparallel pre-existing


joints after different freeze-thaw cycles.

The Influence of FTCs on Microstructure of Sandstone Samples

To explore the influence of FTCs on the microstructure of sandstone samples with pre-
existing joints, the micro-cracks distribution after FTCs is given as shown in Figure 6. There is
no micro-crack in the sample before 45 FTCs. After 45 FTCs, we can observe some micro-
cracks appear on the top and right surfaces of the sample as shown in Figure 6(a). This

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 259

phenomenon can be explained by the fact that the particles close to the center of sample are
subjected to more constraints compared with the particles near the surfaces of the sample. When
the frost heaving force caused by FTCs acts on surrounding rock particles, the contacts close to
the boundary are easier to break compared with the contacts near the center of sample (Huang et
al. 2022). Therefore, with the increase of FTCs, the micro-cracks tend to appear on the surfaces
of sample at first (Zhu et al. 2021).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 6. Micro-cracks distribution after FTCs: (a) 45 FTCs; (b) 60 FTCs.

As FTCs increase, more and more micro-cracks appeared as shown in Figure 6(b). We can
observe some micro-cracks at the tip of the left pre-existing joint. The damage caused by FTCs
deteriorates the microstructure of samples, which explains why the mechanical properties (UCS
and Young’s modulus) of sandstone samples decrease with the increase of FTCs.

(a) 7.24 MPa (b) 13.08 MPa (c) 22.27 MPa (d) 31.28 MPa

Figure 7. Crack propagation and coalescence process of sandstone sample subjected to 60


FTCs (red lines = microcracks; green lines = pre-existing joints).

Crack Propagation and Coalescence of Sandstone Sample Subjected to 60 FTCs

Figures 7 and 8 show the crack propagation and coalescence process of sandstone sample
subjected to 60 FTCs in numerical and experimental uniaxial compressive tests at different axial
stresses. When the axial stress increases to 7.24 MPa as shown in Figure 7(a), a crack initiated

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 260

from the middle of the right pre-existing joint and then propagated along the loading direction
(Figure 7(b)), corresponding to the crack 1a shown in Niu et al. (2019). When the axial stress
increases to 22.27 MPa, some micro-cracks appeared around the inner tips of pre-existing joints,
coalesced with the micro-cracks appeared in FTCs stage. The damage zone formed in the rock
bridge area was same with the experimental results. With the increasing of axial stress, more and
more micro-cracks coalesced with others which formed the splitting failure from the macro
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

perspective.

CONCLUSIONS

This study investigated the influence of freeze-thaw cycles on the mechanical properties of
sandstone sample with two unparallel pre-existing joints by developing an elastoplastic parallel
bond model in PFC2D. The phase transition from water to ice deteriorated the structure of rock
sample and produces micro-cracks in the sample, resulting in the degraded mechanical properties
(UCS and Young’s modulus) of rock. The UCS and Young’s modulus decreased with the
increase of freeze-thaw cycles. With more FTCs, sandstone samples showed more nonlinear
deformation. These trends have good agreements with published experimental data, which
validated that the damage caused by FTCs can be captured by the developed elastoplastic parallel
bond model.
Crack propagation and coalescence processes of sandstone samples subjected to 60 FTCs in
numerical and experimental tests illustrated the development of failure. The micro-cracks
appeared around the pre-existing joints and coalesced with the micro-cracks caused by FTCs.
The numerical results show the formation of cracks and damage zone observed in experiments
from the micro perspective. This research enhances the understanding of the flawed rock
degradation process under FTCs.

REFERENCES

Huang, C., Yang, W., Duan, K., Fang, L., Wang, L., and Bo, C. (2019). “Mechanical behaviors
of the brittle rock-like specimens with multi-non-persistent joints under uniaxial
compression.” Constr Build Mater, 220, 426-443.
Huang, C., Zhu, C., Ma, Y., and Aluthgun Hewage, S. (2022). “Investigating Mechanical
Behaviors of Rocks Under Freeze–Thaw Cycles Using Discrete Element Method.” Rock
Mech Rock Eng.
Huang, D., and Zhu, T. T. (2018). “Experimental and numerical study on the strength and hybrid
fracture of sandstone under tension-shear stress.” Eng Fract Mech, 200, 387-400.
Khanlari, G., Sahamieh, R. Z., and Abdilor, Y. (2014). “The effect of freeze–thaw cycles on
physical and mechanical properties of Upper Red Formation sandstones, central part of Iran.”
Arabian J Geosci, 8(8), 5991-6001.
Lu, Y., Li, X., and Chan, A. (2019). “Damage constitutive model of single flaw sandstone under
freeze-thaw and load.” Cold Reg Sci Technol, 159, 20-28.
Niu, Y., Zhou, X. P., Zhang, J. Z., and Qian, Q. H. (2019). “Experimental study on crack
coalescence behavior of double unparallel fissure-contained sandstone specimens subjected
to freeze-thaw cycles under uniaxial compression.” Cold Reg Sci Technol, 158, 166-181.
Wang, Y., Han, J. Q., and Li, C. H. (2020). “Acoustic emission and CT investigation on fracture
evolution of granite containing two flaws subjected to freeze–thaw and cyclic uniaxial
increasing-amplitude loading conditions.” Constr Build Mater, 260.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 261

Zhou, X. P., Niu, Y., Zhang, J. Z., Shen, X. C., Zheng, Y., and Berto, F. (2019). “Experimental
study on effects of freeze‐thaw fatigue damage on the cracking behaviors of sandstone
containing two unparallel fissures.” Fatigue & Fracture of Engineering Materials &
Structures, 42(6), 1322-1340.
Zhu, T., Chen, J., Huang, D., Luo, Y., Li, Y., and Xu, L. (2021). “A DEM-Based Approach for
Modeling the Damage of Rock Under Freeze–Thaw Cycles.” Rock Mech Rock Eng(54),
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2843-2858.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 262

Development of a Temperature-Controlled Direct Shear Box for Frozen Samples

Hossein Emami Ahari, S.M.ASCE1; and Beena Ajmera, Ph.D., P.E., M.ASCE2
1
Graduate Student, Dept. of Civil, Construction, and Environmental Engineering, Ames, IA.
Email: hemami@iastate.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Assistant Professor, Dept. of Civil, Construction, and Environmental Engineering, Iowa State
Univ., Ames, IA. Email: bajmera@iastate.edu

ABSTRACT

Global surface temperatures rose by 1°C in 2020 relative to the average temperatures
between 1951 and 1980. A reduction in the extent of the almost 24% of the Northern
Hemisphere covered by permafrost will impact the stability of new and existing civil
infrastructure in the area. Frozen soil is a multi-phase system consisting of ice, unfrozen water,
air, and soil particles. Fluctuations in temperature change the proportions of the constituents
significantly impacting the strength of the soil mass. The strength behavior of fine-grained soils
will differ substantially from that of frictional soils since fine-grained soils can have high ice
contents due to their high-water absorption capacities. However, the lack of commercially
available equipment has deterred research efforts in understanding the impact of climate change
on the shear strength properties of fine-grained soils. This paper describes the development of a
new direct shear box that allows for temperature-controlled testing. Modifications to the
temperature-controlled shear box as deemed necessary from validation tests are also described.
For example, comparisons of the temperature measurements within the soil sample to the chilled
glycol circulating in the modified direct shear box indicated that additional insulation had to be
provided to the apparatus. The freezing mechanism applied to the soil sample affected the
distribution of ice within the pore spaces and illustrated the need to freeze the sample from all
directions. This was incorporated into the modifications that were designed in this study. Shear
strength parameters from the newly designed temperature-controlled direct shear box matched
well with those from the traditional shear box.

INTRODUCTION

Global climate has warmed at the rate of 0.08°C per decade since 1880 relative to the
average temperature in the century from 1901 to 2000. In the last forty years, this rate has more
than doubled to 0.18°C per decade (NOAA 2019). Moreover, the ten warmest years on record
have occurred over the last eighteen years (NOAA 2019 and NASA 2020). Due to the emission
of carbon dioxide and other greenhouse gas emissions, the earth is expected to experience
increasingly warmer climates over the next few decades (NASA 2019). The warming of the
global climate will induce changes in the surface temperature with greater affects in Artic
regions as a result of the Arctic amplification phenomena (Serreze and Francis 2006).
Permafrost is ground that remains at subzero temperatures for at least two consecutive years.
It constitutes about 25% and 75% of the land area of the Northern Hemisphere and Alaska,
respectively (Osterkamp and Lachenbruch 1990). Temperature data from drilled holes in the
permafrost of the Arctic Coastal Plain and the Brooks Range in Alaska showed that global
warming caused a 2°C to 4°C increase in the mean surface temperature over the last century

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 263

(Osterkamp and Lachenbruch 1990). Since the temperature in some regions of the permafrost is
only a few degrees below the melting point, increases in the mean surface temperature as a result
of climate change will induce extensive thawing of the underlying soils. For example, Lawrence
and Slater (2005) showed that between 1900 and 2000, about 25% of the permafrost region was
lost. More than 58% of the Canadian permafrost will be lost as a result of a 5°C rise in the mean
annual air temperature (Buteau et al. 2010).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Loss of the permafrost will have substantial geotechnical effects. One effect is the triggering
of slope movements as a result of the thawing. For example, a thaw-induced landslide in the
permafrost region near the Thunder River has been estimated to threaten the Mackenzie Valley
pipeline in Canada, valued at $7 billion. This pipeline transfers natural gas to markets in
Southern Canada and the United States (Singhroy 2009). Another example of the geotechnical
consequences resulting from the thaw of permafrost is the loss of bearing capacity. In one of the
largest cities in Northwest Siberia, Nadym, bearing capacities for foundations in 2000 appeared
to be 33% lower than those calculated in 1960 (Streletskiy et al. 2012). It was further found that
the bearing capacity will reduce by an additional 52% by 2040.
Since changes in the surface temperature will affect various geotechnical engineering
properties, many past studies have focused on evaluating the fluctuations in these properties with
an increase in temperature. For example, Anderson et al. (1995) found that a rise in the
temperature from -20°C to -10°C would cause a reduction in the yield stress in Manchester fine
sand. Mixtures of ice and fine sand were subjected to direct shear tests at normal stress between
200 kPa and 400 kPa at temperatures ranging from -7°C to -3°C in a study by Yasufuku et al.
(2003). Their results indicated that the peak shear strength decreased as the temperature was
increased. Arenson and Springman (2005a) drew similar conclusions regarding the peak shear
strength of rock glaciers and artificially frozen samples from the results of consolidated
undrained (CU) triaxial tests conducted at temperatures between -4°C and -1°C. In their study,
residual shear strengths were also found to decrease with an increase in the temperature (Arenson
and Springman 2005a). Such reductions in shear strength as a result of increases in the
temperature may be attributed to the loss of suction as ice within the soil mass melts increasing
the unfrozen water content of the sample.
A detailed review of the prior publications related to the shear strength of soils that were
once frozen (Andersland and AlNouri 1970; Parameswaran 1980; Bragg and Andersland 1981;
Ladanyi 1972, 1985; Watson et al. 1973; Sayles and Haines 1974; Zhu and Carbee 1984; Hivon
and Sego 1995; Christ and Kim 2009; Buteau et al. 2010; Yang et al. 2015; Anderson et al. 1995;
Yasufuku et al. 2003; Arenson and Springman 2005a, 2005b; Arenson 2002; Frankenstein and
Shoop 2015; Yamamoto and Springman 2014, 2019; Yugui et al. 2016; among others) highlights
a focus of past research in understanding the behavior of coarse-grained (frictional) soils.
However, frictional soils will be less prone to changes in shear strength as a result of the thawing
process in comparison to fine-grained (cohesive) materials. In particular, the greater variations in
the shear strength of fine-grained soils can be attributed to higher specific areas allowing these
materials to observe large amounts of water. From the few studies that have examined the shear
strength of fine-grained soils, it is evident that the effect of temperature is still the prime focus
(Agregaard and Ingeman-Nielsen 2012; Fernandez Santoyo et al. 2021). However, there are
many other factors that will influence the shear strength of soils, and especially fine-grained
soils, that have not been extensively studied. This gap may be related to the difficulties in
conducting temperature-controlled tests and the lack of commercially available equipment to
obtain high-quality results. This paper describes the development of a temperature-controlled

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 264

direct shear box. A prototype of this box was built and tested. Preliminary results from the
testing and validation efforts undertaken by the authors are also presented in this paper.

DEVELOPMENT OF TEMPERATURE-CONTROLLED DIRECT SHEAR BOX

The configuration of the traditional shear box does not allow for accurate testing of samples
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

under temperature-controlled settings. Such testing conditions are of prime importance as


changes in the climate impact the ground temperatures require geotechnical engineers to quantify
the shear strength at different temperatures. To date, such testing has occurred in specialized
triaxial devices (for example, in Arenson and Springman 2005a) or using traditional testing
equipment placed in temperature-controlled rooms (for example, in Yasufuku et al. 2003). In this
study, a new direct shear box is designed that modifies traditional direct shear devices to
facilitate temperature-controlled testing. Additionally, the designed direct shear box meets the
guidelines set forth in ASTM D3080/D3080M to ensure its usability with traditional direct shear
devices allowing the strengths obtained from this device to be applicable to geotechnical design
and evaluation standards and codes.
The newly developed temperature-controlled shear box consists of two halves (a top and
bottom) and a direct shear cap, as shown in Figure 1. Each of these components is hollow to
allow for the circulation of a fluid that will impose the desired temperature on all dimensions of
the sample being tested. To obtain a frozen sample, all the inflow and outflow valves of the
temperature-controlled direct shear box were connected to a chiller (cryostat) which circulates
glycol inside each component at the desired temperature. A schematic of the test set-up is shown
in Figure 2. To maintain the temperature of the sample within the direct shear box, the box was
insulated, as illustrated in Figure 3. However, modifications had to be made to the insulation in
order to achieve the desired temperature control, as will be detailed in the next section.

Figure 1. (a) Components of Newly Developed Temperature-Controlled Direct Shear Box


and (b) Assembled Direct Shear Box.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 265

VERIFICATION EFFORTS

Temperature Control. Figure 4 illustrates the temperature of a soil sample inside the
temperature-controlled direct shear box. The temperature of the glycol in the chiller that was
circulating in the newly developed direct box is also shown in the figure. As shown in Figure 4,
the soil temperature reached -4°C after about two hours. After this point, the temperature
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

remained unchanged for the approximately eight hours over which the new direct shear box was
tested. During this time, the chiller was circulating glycol chilled to a temperature of -10°C. This
indicated that a temperature loss of 6°C occurred with this insulation scheme. To reduce the
temperature loss, the entire direct shear set-up was insulated using high-density polystyrene and
insulation foam, as shown in Figure 2. With these modifications to the insulation, temperature
loss between the soil and the circulating glycol were reduced to approximately 1°C.
Uniformity of Ice Lenses in Soil Samples. The mechanism by which the sample is frozen
affects the formation of ice lenses within the soil mass. When a soil sample is frozen uni-
directionally from the top downward, a temperature gradient is produced across the sample
inducing cryogenic suction. This suction will cause water in the voids to migrate towards the
freezing front, as shown in Figure 5. As a result, when the water freezes, the ice will not be
uniformly distributed throughout the soil mass. Rather, ice lenses will be produced, as shown in
Figures 5c and 5d.

Figure 2. Schematic of Direct Shear System for Temperature-Controlled Tests.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 266
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Foam Insulation Placed Around Temperature-Controlled Direct Shear Box.

Figure 5 clearly illustrates that samples must be frozen from all directions to avoid the
migration of water resulting in the development of non-uniform ice distribution within the
sample when it freezes. This observation was a key consideration during the development of the
direct shear box for temperature-controlled testing. Hence, it was necessary to circulate chilled
glycol across both halves of the modified direct shear box as well as the top platen, as described
previously. Figure 6 contains pictures of a sample that was subjected to uniform freezing from all
directions in the modified direct shear box. It is clear that the ice content is uniformly distributed
across the height of the sample (Figure 6a) and across its diameter (Figure 6b).

Figure 4. Temperature of Soil within Modified Shear Box and Circulated Glycol.

Shear Strength Parameters of Ottawa Sand. Verification of the measured shear strength
parameters were performed using 10-30 Ottawa sand samples. Dry samples were compacted into
the standard (traditional) direct shear box as well as the developed temperature-controlled direct
shear box. The density of the sample was held constant at 17 kN/m3. Normal stresses of 69 kPa,
103 kPa, and 138 kPa, were applied on separate samples. Once the samples had consolidated
under these pressures, they were sheared at a rate of 1.02 mm/min. The samples were sheared
until the peak strength was obtained or for a maximum shear strain of 25%. Each test was
repeated at least three times using separate samples.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 267
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5. (a) Uni-directional Freezing of Sample from Top, (b) Schematic Illustrating
Water Migration, (c) Picture of Uni-directional Sample Frozen from Top, and (d) Dissected
Sample Illustrating Non-Uniform Ice Content.

Figure 6. Ice Content Distribution across (a) Height and (b) Diameter of Sample Frozen
from All Directions.

Figure 7 contains an example of the variation in shear stress as a function of the horizontal
stress at a normal stress of 103 kPa from both the traditional shear box and the temperature-

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 268

controlled shear box. As seen from this figure, the differences between the curves can be
considered negligible. Similar results were obtained at all of the normal stresses tested in this
study. This indicates the temperature-controlled shear box was able to properly reproduce soil
behavior.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 7. Horizontal Displacement versus Shear Stress at Normal Stress of 103.4 kPa.

The shear envelopes obtained from the verification tests conducted are shown in Figure 8.
The cohesion intercept obtained from tests conducted using both the traditional and temperature-
controlled shear boxes were zero, as expected. The friction angle from the traditional direct shear
box was found to be 35°, only 0.3° higher than the friction angle obtained from the tests
conducted using the temperature-controlled shear box. This difference is considered negligible
and may be attributed to slight variations in the sample preparation process or consolidation
stages.

=° =°

Figure 8. Shear Envelopes for Ottawa Sand Samples Tested using (left) a Traditional and
(right) the Temperature-Controlled Shear Box.

Feasibility of Temperature-Controlled Direct Shear Testing. In order to assess the


feasibility of temperature-controlled direct shear testing, the shear strength of soil was measured

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 269

at different temperatures and compared to that of the standard direct shear box at room
temperature. Soil specimens were prepared using a commercially available kaolinite clay that
was mixed with an initial moisture content of the liquid limit of 73 (Tiwari and Ajmera 2011)
following which the sample was allowed to hydrate for at least 24 hours. The resulting slurry
was consolidated to a stress of 50 kPa.
For the temperature-controlled testing, after the completion of the primary consolidation, as
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

monitored by real-time logarithm of time versus displacement curves, a chiller was connected to
the direct shear box to circulate glycol chilled at the desired temperature. To ensure a uniform
distribution of temperature across the frozen sample, the soil specimens were allowed to freeze
for a period of at least twelve hours before they were sheared at a rate of 0.001 mm/min.
Shearing continued until the samples reached their peak strength or for a maximum shear strain
of 25%.
In the traditional direct shear test, the sample was sheared immediately after the primary
consolidation was complete. The shearing rates and termination criteria were the same as for the
temperature-controlled testing. The room temperature during this test was 20°C.
Figure 9 demonstrates the shear stress with horizontal displacement obtained from the
temperature-controlled and standard direct shear boxed. This figure illustrates the significant
differences in the strengths of the frozen and unfrozen kaolinite samples. For the scope of this
paper, this figure demonstrates that the temperature-controlled direct shear box developed in this
study contains the functionality to conduct temperature-controlled testing and is able to capture
the differences in shear strength of samples at various temperatures.

Figure 9. Comparison of Horizontal Displacement versus Shear Stress Curves for Kaolinite
at a Normal Stress of 50 kPa from a Standard Direct Shear Box at Temperature of +20℃
and from a Temperature-Controlled Direct Shear Box at Temperatures of -10℃, -8℃, and
-6℃.

CONCLUSIONS

The rise in global surface temperatures and the geotechnical issues that arise highlight the
need to examine the fluctuations in soil properties as a function of temperature and various other

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 270

parameters associated with frozen soils. However, past research in this area has been performed
using equipment that needed substantial modifications in order to perform temperature-
controlled testing or through the use of equipment that was placed in a temperature-controlled
room. In this study, simple modifications to a traditional direct shear box in order to perform
temperature-controlled tests were described.
Comparison of the soil temperature inside of the temperature-controlled direct shear box with
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the temperature of the chilled circulating glycol indicated a loss of 6°C. As a result, the
insulation mechanism of the temperature-controlled direct box (use of insulating foam on the
shear box) needed modification. High-density polystyrene and insulation foam were used to
insulate the apparatus reducing the temperature loss to 1°C.
Freezing of the sample uni-directionally from the top results in the migration of the water
towards the freezing front. This results in a non-uniform distribution of ice when the water
freezes with more ice present in the top half of the sample and a nearly dry material in the
bottom half of the sample. Ice lenses appear to concentrate at freezing front. Freezing of the
sample from all directions, as possible with the temperature-controlled shear box developed in
this study, results in the uniform distribution of ice along the height and diameter of the sample.
Verification tests were performed using Ottawa sand samples. The results illustrated that the
variation in shear stress as a function of the horizontal displacement is negligible. Similarly, little
differences were found between the shear strength parameters (that is, the cohesion intercept and
friction angle) of the samples tested. This indicated that the temperature-controlled direct shear
box will properly capture soil behavior.
The functional capability of the temperature-controlled direct shear box was assessed using
kaolinite clay samples. The result showed the significantly higher shear strength of frozen
kaolinite samples in comparison to the unfrozen sample tested. The results confirmed the ability
of the newly developed temperature-controlled direct shear box to perform direct shear tests at
controlled temperatures.

REFERENCES

Agergaard, F. A., and Ingeman-Nielsen, T. (2012). “Development of Bearing Capacity of Fine-


Grained Permafrost Deposits in Western Greenland Urban Areas Subject to Soil Temperature
Changes,” Proceedings of Cold Regions Engineering 2012, 82-92.
Andersen, G. R., Swan, C. W., Ladd, C. C., and Germaine, J. T. (1995). “Small-strain Behavior
of Frozen Sand in Triaxial Compression,” Canadian Geotechnical Journal, 32(3), 428-451.
Andersland, O. B., and AlNouri, I. (1970). “Time-dependent Strength Behavior of Frozen Soils,”
Journal of the Soil Mechanics and Foundations Division, 96(4), 1249–1265.
Arenson, L. U. (2002). Unstable Alpine Permafrost: A Potentially Important Natural Hazard,
Institute for Geotechnical Engineering, Swiss Federal Institute of Technology ETH Zurich.
Arenson, L. U., and Springman, S. M. (2005a). “Triaxial Constant Stress and Constant Strain
Rate Tests on Ice-rich Permafrost Samples,” Canadian Geotechnical Journal, 42(2), 412-
430.
Arenson, L. U., and Springman, S. M. (2005b). “Mathematical Descriptions for the Behaviour of
Ice-rich Frozen Soils at Temperatures Close to 0°C,” Canadian Geotechnical Journal, 42(2),
431-442.
ASTM. ASTM D3080/D3080M. (2011). Standard Test Method for Direct Shear Test of Soils
under Consolidated Drained Conditions, ASTM International.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 271

Bragg, R. A., and Andersland, O. B. (1981). “Strain Rate, Temperature, and Sample Size Effects
on Compression and Tensile Properties of Frozen Sand,” Engineering Geology, 18(1–4), 35–
46.
Buteau, S., Fortier, R., and Allard, M. (2010). “Permafrost Weakening as a Potential Impact of
Climatic Warming,” Journal of Cold Regions Engineering, 24(1), 1-18.
Christ, M., and Kim, Y.-C. (2009). “Experimental Study on the Physical-mechanical Properties
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

of Frozen Silt.” KSCE Journal of Civil Engineering, 13(5), 317–324.


Fernandez Santoyo, S., Tom, J. G., and Baser, T. (2021). “Impact of Subsurface Warming on the
Capacity of Helical Piles Installed in Permafrost Layers,” Proceedings of the International
Foundations Conference and Equipment Expo 2021, 323, 239-248.
Frankenstein, S., and Shoop, S. A. (2015). “Characterization of the Physical Properties of Frozen
Soil under Varying Ice Content,” Proceedings of Cold Regions Engineering 2015, 13–24.
Hivon, E. G., and Sego, D. C. (1995). “Strength of Frozen Saline Soils,” Canadian Geotechnical
Journal, 32(2), 336–354.
Ladanyi, B. (1972). “An Engineering Theory of Creep of Frozen Soils,” Canadian Geotechnical
Journal, 9(1), 63-80.
Ladanyi, B. (1985). “Use of the Cone Penetration Test for the Design of Piles in Permafrost.”
Journal of Engineering Resources Technology, 107(2), 183–187.
Lawrence, D. M., and Slater, A. G. (2005). “A Projection of Severe Near-surface Permafrost
Degradation during the 21st Century,” Geophysical Research Letters, 32(24), L24401.
NASA Goddard Institute for Space Studies. (2020). “Global Temperature,” NASA Climate
Change Vital Signs of the Planet. <https://climate.nasa.gov/>.
NOAA. (2019). “State of the Climate: Annual 2018 Global Climate Report,” National Centers
for Environmental Information of the National Oceanic and Atmospheric Administration.
<www.ncdc.noaa.gov/sotc/global/201813>.
Osterkamp, T. E., and Lachenbruch, A. H. (1990). “Thermal Regime of Permafrost in Alaska
and Predicted Global Warming,” Journal of Cold Regions Engineering, 4(1), 38-42.
Parameswaran, V. R. (1980). “Deformation Behaviour and Strength of Frozen Sand,” Canadian
Geotechnical Journal, 17(1), 74–88.
Sayles, F. H., and Haines, D. (1974). Creep of Frozen Silt and Clay, U.S. Army Corps of
Engineers Cold Regions Research and Engineering Laboratory.
Serreze, M. C., and Francis, J. A. (2006). “The Arctic Amplification Debate,” Climatic Change,
76(3), 241-264.
Singhroy, V. (2009). “Satellite Remote Sensing Applications for Landslide Detection and
Monitoring,” Landslides – Disaster Risk Reduction, 143-158.
Smith, M. (1990). “Potential Responses of Permafrost to Climatic Change,” Journal of Cold
Regions Engineering, 4(1), 29-37.
Streletskiy, D. A., Shiklomanov, N. I., and Nelson, F. E. (2012). “Permafrost, Infrastructure, and
Climate Change: A GIS-Based Landscape Approach to Geotechnical Modeling,” Arctic,
Antarctic, and Alpine Research, 44(3), 368-380.
Tiwari, B., and Ajmera, B. (2011). “A new correlation relating the shear strength of reconstituted
soil to the proportions of clay minerals and plasticity characteristics” Applied Clay Science,
53(1), 48-57.
Watson, G. H., Slusarchuk, W. A., and Rowley, R. K. (1973). “Determination of Some Fozen
and Thawed Properties of Permafrost Soils,” Canadian Geotechnical Journal, 10(4), 592–
606.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 272

Yamamoto, Y., and Springman, S. (2014). “Axial Compression Stress Path Tests on Artificial
Frozen Soil Samples in a Triaxial Device at Temperatures Just Below 0°C,” Canadian
Geotechnical Journal, 51, 1178–1195.
Yamamoto, Y., and Springman, S. M. (2019). “Triaxial Stress Path Tests on Artificially Prepared
Analogue Alpine Permafrost Soil,” Canadian Geotechnical Journal, 56, 1448-1460.
Yasufuku, N., Springman, S. M., Arenson, L. U., and Ramholt, T. (2003). “Stress-dilatancy
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Behaviour of Frozen Sand in Direct Shear,” Proceedings of the 8th International Conference
on Permafrost, 1253-1258.
Yugui, Y., Feng, G., Yuanming, L., and Hongmei, C. (2016). “Experimental and Theoretical
Investigations on the Mechanical Behavior of Frozen Silt,” Cold Regions Science and
Technology, 130, 59–65.
Zhu, Y., and Carbee, D. L. (1984). “Uniaxial Compressive Strength of Frozen Silt under
Constant Deformation Rates,” Cold Regions Science and Technology, 9(1), 3–15.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 273

Modified Hyperbolic Model for Dynamic Properties of Peaty Organic Soils

Pengfei Wang, Ph.D., M.ASCE1; Tristan E. Buckreis, S.M.ASCE2;


Scott J. Brandenberg, Ph.D., P.E., M.ASCE3; and Jonathan P. Stewart, Ph.D., P.E., F.ASCE4
1
Assistant Professor, Dept. of Civil and Environmental Engineering, Old Dominion Univ.,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Norfolk, VA; formerly, Postdoctoral Scholar, Dept. of Civil and Environmental Engineering,
Univ. of California, Los Angeles, CA (corresponding author). Email: p1wang@odu.edu
2
Graduate Student, Dept. of Civil and Environmental Engineering, Univ. of California,
Los Angeles, CA. Email: tristanbuckreis@ucla.edu
3
Professor, Dept. of Civil and Environmental Engineering, Univ. of California, Los Angeles, CA.
Email: sjbrandenberg@ucla.edu
4
Professor, Dept. of Civil and Environmental Engineering, Univ. of California, Los Angeles, CA.
Email: jstewart@seas.ucla.edu

ABSTRACT

Modulus reduction and damping versus shear strain relationships (hereafter MRD) are
essential for nonlinear ground response analyses of soil sites. For sites containing thick deposits
of peat, large shear strains that produce pronounced nonlinearity are expected in seismically
active regions. Currently, available MRD models for peat have strongly divergent characteristics,
particularly with regard to the influence of confining stress. We assembled a large database of
test data from literature and proposed a hyperbolic model, similar in form to widely used models
for non-organic soils, for application to peat. We related the critical model parameters to stress
history-modified effective stress, finding no strong dependence on organic content (OC) and peat
composition (fibrous structure) over the parametric range of the database (OC >30%). The new
model captures both the weak and strong dependencies of MRD behaviors with effective stress
from literature, with weak dependence occurring in normally consolidated peats for 𝜎𝑣, /𝑝𝑎 > 0.2
(where 𝜎𝑣, is vertical effective stress and 𝑝𝑎 is atmospheric pressure) and stronger dependence at
smaller stresses. We present here a model for OC >30% conditions and its associated aleatory
variability, which was established from residuals derived from test results. Compared to previous
peat MRD models, the proposed model has more easily interpretable parameters and reduces the
model bias and uncertainty.

INTRODUCTION

The modulus reduction and damping versus shear strain relationship (hereafter MRD), of
peaty organic soils, is a significant component of ground response analysis for regions with thick
deposits of peat. For example, in the Sacramento-San Joaquin Delta, California, a great deal of
critical infrastructure for water and gas distribution is underlain by thick deposits of highly
organic peaty soils. As such, the MRD behavior of this peat is critical for seismic hazard
characterization in this region.
Previously-published MRD relationships for peat indicate significant different behaviors,
particularly in relation to the impact of effective confining stress on the degree of nonlinearity.
Some experimental studies show a large effect of confining pressure (e.g., Kramer 2000 for
Mercer Slough peat in Washington and Wehling et al. 2003 for Sherman Island peat), whereas
Kishida et al. (2009; hereafter Kea09) indicate such dependencies are relatively weak. Moreover,

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 274

Kea09 uses a complicated functional form that does not allow model parameters to be clearly
understood and compared to data trends or other models. To resolve these issues, we present a
new model for peat MRD. We first describe a newly assembled global database from literature.
We then present a hyperbolic model for modulus reduction that is modified from prior
hyperbolic relations for inorganic soils by Darendeli (2001) and Menq (2003). Critical elements
within the framework are pseudo-reference strain 𝛾𝑟 (shear strain where modulus reduction
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

factor is 0.5) and minimum damping 𝐷min (damping at small strain). We relate these parameters
to stress history-modified effective stress and investigate the dependencies on organic content
(OC) and peat composition (fibrous structure). Lastly, we utilize the entire shear modulus
reduction and damping curve data to fit a few remaining coefficients to develop the complete
MRD model and derive standard deviations.

DATABASE

We have assembled the results of cyclic tests performed on peat samples worldwide to
measure MRD soil behavior from previous literature. The test results are grouped by region
(Washington, Japan, Greece, Netherlands, Delta, and New York). References and major features
of the data from each region are provided in Table 1. The data collectively covers a broad range
of overburden pressures from 1.5 to 400 kPa and organic contents (OC) from 14 to 87 %. While
most of the peat specimens were normally consolidated or lightly overconsolidated for dynamic
testing, data from the Netherlands has overconsolidation ratios (OCRs) up to 3.8. The peat
specimens also show diverse fibrous structures, from fibrous to decomposed.

MODEL DEVELOPMENT

We use the database to fit a new MRD model for peat. The model has two critical
parameters, pseudo-reference shear strain and minimum damping, which are related to
predictors. The effects of additional parameters and aleatory variability are then estimated.

Model for pseudo-reference strain

The pseudo-reference shear strain, 𝛾𝑟 , is the shear strain at which the shear modulus, 𝐺(𝛾), is
the half of maximum shear modulus (𝐺max ), such that 𝐺(𝛾𝑟 )⁄𝐺𝑚𝑎𝑥 = 0.5. The hyperbolic model
uses 𝛾𝑟 as defined in Eq. 1, where a is a model curvature parameter.

𝐺(𝛾) 1
= (1)
𝐺max 1 + ( 𝛾 )𝑎
𝛾𝑟

Using the developed database, we derived values of 𝛾𝑟 from peat MRD tests. In Fig. 1, we
plot interpreted pseudo-reference shear strains, 𝛾𝑟 , versus normalized overburden pressure,
𝜎𝑣, ⁄𝑝𝑎 . The data for the five regions in Table 1 are marked with different colors. The data in Fig.
1 suggest a strong impact of 𝜎𝑣, on 𝛾𝑟 for 𝜎𝑣, ⁄𝑝𝑎 < 0.3, which weakens for larger 𝜎𝑣, . For very
small overburden pressures, 𝜎𝑣, ⁄𝑝𝑎 < 0.1, there is a large difference between the two
Netherlands data points (purple dots, 𝛾𝑟 ≈ 1%) and most of the other test data, mainly from
Washington (𝛾𝑟 < 0.1%). The Washington and Netherlands data have similar OCs, but different
stress histories (OCR=1.4 – 3.8 for Netherlands data versus OCR=1 for Washington data).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 275

Table 1. Summary of peat MRD test data

Fibrous
Region Locations References # Tests 𝝈,𝒗 (𝐤𝐏𝐚) OC (%) OCR Tests*
Structure
Mercer
Washington Kramer (2000) Fibrous 14 1.5 – 30 72 – 80 1 RC, TX
Slough
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Tokimatsu and
Japan Ojiya Unknown 2 29 – 39 62 1 TX
Sekiguchi (2006)
Philippi
Decomposed 1 370 48 1.8 RC
peatland Kallioglou et al.
Greece
Ptolemaida- (2008)
Fibrous 1 400 62 1.3 RC
Kozani
Zwanenburg et al.
Netherlands Groningen Fibrous 3 6.2 – 28 80 – 87 1.4 – 3.8 RC, DSS
(2020)
Sherman Wehling et al.
Fibrous 13 10 – 86 21 – 52 1 TX
Island (2003)
Boulanger et al. 8 128 – 200 44 – 63 1 TX
Sherman (1998) &
Fibrous
Island Arulnathan et al. 2 66 63 – 65 2 TX
(2000)
Sacramento- Montezuma 7 16 – 272 15 – 61 1 TX
Kishida et al.
San Joaquin Decomposed TX, RC,
Slough (2009b) 1 202 38 2
Delta, TS
California TX, RC,
3 68 – 138 17 – 34 1
Kishida et al. TS
Clifton Court Fibrous
(2009b) TX, RC,
1 34 20 2
TS
6 55.2 – 220.9 14 – 35 1 RC, TS
Stokoe (2003,
Clifton Court Unknown 1 34.5 30 2 RC, TS
personal comm.)
3 17.3 – 55.2 14 – 35 4 RC, TS
Queensboro
New York Stokoe et al. (1994) Unknown 2 76 35 – 63 1 RC, TS
Bridge
* Note, RC, TX, TS, and DSS represent Resonant Column, Cyclic Triaxial, Cyclic Torsion Shear, and
Cyclic Direct Simple Shear Tests, respectively

Figure 1. Variation of peat pseudo-reference shear strains with overburden stress

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 276
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Variation of peat pseudo-reference shear strain with OCR-adjusted overburden


stress

To evaluate whether stress history may explain the different levels of nonlinearity in the
Washington and Netherlands samples, we plot in Fig. 2 the same data as in Fig. 1 but with an
OCR-adjusted normalized overburden pressure, (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 , on the abscissa. We assume
the exponent parameter 𝑚 = 0.93 based on the OCR effect implied by the Darendeli (2001) 𝛾𝑟
model (specifically, 𝛾𝑟 = 𝜙3 ⁄𝜙4 , where the 𝜙 terms are model coefficients with values 𝜙3 =
0.3246 and 𝜙4 = 0.3483). As shown in Fig. 2, the use of the OCR-adjusted normalized
overburden pressure shifts the Netherlands data to the right and provides a more consistent data
trend with reduced scatter.
We develop a model to characterize the first-order dependence of 𝛾𝑟 on overburden stress
and OCR, with modeling of OC effects deferred to a later stage of model development. The
model in this form is similar to a prior peat MRD model by Rodriguez-Marek et al. (2017)
(hereafter Rea17). The model for 𝛾𝑟 is given in Eq. (2),
𝑝𝑎
𝜎𝑣, ∗ OCR𝑚 + 𝑔3
log(𝛾𝑟 ) = 𝑔1 + 𝑔2 ∗ log ( ) (2)
𝑔3

where 𝛾𝑟 is in percent, 𝑔1 is the asymptotic log(𝛾𝑟 ) as 𝜎𝑣, increases to infinity, 𝑔2 is the gradient
𝑝𝑎 ,
of log(𝛾𝑟 ) increase with respect to log (𝜎, ∗OCR 𝑚 ) for small 𝜎𝑣 (a larger value indicates a steeper
𝑣
gradient), and 𝑔3 is related to the transition of OCR-adjusted normalized stress beyond which
log(𝛾𝑟 ) is relatively constant. When 𝜎𝑣, ∗ OCR𝑚 is greater than 𝑝𝑎 /𝑔3 , log(𝛾𝑟 ) increases slowly
with increasing 𝜎𝑣, ∗ OCR𝑚 until reaching a plateau.
Nonlinear regression based on the maximum log-likelihood method (equal weight to all data
points) provides fitted parameters, 𝑔1 = 0.4620, 𝑔2 = −2.1197, and 𝑔3 = 4.3153. The
resulting fit is shown in Fig. 3. These regression coefficients indicate that the high-stress
𝑝
saturation of 𝛾𝑟 occurs at 𝜎𝑣, ∗ OCR𝑚 > 𝑔𝑎 ≈ 23 kPa, and that the asymptotic reference strain for
3

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 277

high-stress conditions is 𝛾𝑟 = exp(𝑔1 ) ≈ 1.59%. Model residuals are shown in Fig. 4 with
respect to the model’s independent variable (OCR-adjusted normalized overburden) and OC. The
residuals do not exhibit an appreciable trend with either parameter.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Fit of pseudo-reference shear strain model (Eq. 2) to peat data

Figure 4. Residuals of 𝐥𝐨𝐠(𝜸𝒓 ) data against (𝝈,𝒗 ⁄𝒑𝒂 ) × 𝑶𝑪𝑹𝒎 and OC

Fig. 5 overlays the model on the data (red line), along with predictions of other MRD models
(in black lines) for mineral (non-peat) soils (Darendeli 2001) and peat soils (Kea09 and Rea17).
The Kea09 model was primarily developed using Delta peat data, and it incorporates overburden
pressure and OC as covariates to predict 𝛾𝑟 . The three solid black lines in Fig. 5 from the bottom
to the top are associated with low OC (OC = 20%), intermediate OC (OC = 55%), and high OC
(OC = 80%) predicted by Kea09. The dotted black line in Fig. 5 is the Darendeli01 model
prediction with plasticity index (PI) = 50. The dashed black line in Fig. 5 is from Rea17, which
was developed using Washington, Greece, and some proportion of Delta data (the plot was made
using OCR=1). It assumes that 𝛾𝑟 depends only on 𝜎𝑣, for peat [𝛾𝑟 = 𝑓(𝜎𝑣, )], ignoring OC and
OCR effects.
In Fig. 5, we observe that the Darendeli01 model, which is not intended for organic soil,
underestimates pseudo-reference strain for (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 > 0.1. This is primarily because
the gradient in the Darendeli01 model is too small for peat at low (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 . This
indicates that peat, in general, is more linear than clay. The Rea17 model captures the general
trend of pseudo-reference strain with (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 for (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚  0.1-1, but

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 278

because it does not capture the changes in gradient with (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 , it underestimates and
overestimates pseudo-reference strain for low (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 and high (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚
ranges, respectively. The Kea09 model predicts pseudo-reference strain reasonably well for
(𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 > ~0.2 but overestimates 𝛾𝑟 for (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 < ~0.2.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5. Plot of present model for 𝜸𝒓 and models from literature against peat dataset.

Model for minimum damping

We follow a similar procedure to develop a model for the minimum (low-strain) damping,
𝐷min . Fig. 6 plots the 𝐷min data and shows a fit line derived using the following equation
(adapted from Rea17),

𝛽2
𝜎𝑣′
𝐷min = 𝛽1 ∗ ( ∗ OCR𝑚 ) (3)
𝑝𝑎

The regressed coefficients are 𝛽1 = 2.9800 and 𝛽2 = −0.2332. Residual plots are shown in
Fig. 7 to evaluate model performance and investigate dependencies on OC. The residuals do not
trend markedly with OCR-adjusted normalized overburden stress or OC.

Figure 6. Fit of 𝑫𝐦𝐢𝐧 model (Eq. 3) to peat data

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 279
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 7. Residuals of 𝑫𝐦𝐢𝐧 data against (𝝈,𝒗 ⁄𝒑𝒂 ) × 𝑶𝑪𝑹𝒎 and OC

Fig. 8 compares the present model (red line) and three models from literature (in black lines)
to the data. The three comparison models are Kea09, Rea17, and Darendeli01 (with PI = 50). The
Kea09 model is plotted with three solid black lines corresponding to low OC (OC = 20%),
intermediate OC (OC = 55%), and high OC (OC = 80%) from top to bottom for low (𝜎𝑣, ⁄𝑝𝑎 ) ×
OCR𝑚 . As (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 increases, 𝐷min for three different OC levels get closer and
intersects at (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 ≈ 0.3. The Kea09 curves then cross so that 𝐷min increases with
OC. These trends in the Kea09 model indicate that the decay gradient decreases as OC
increases.
By comparing to data, we see that Kea09 underestimates 𝐷min for (𝜎𝑣, ⁄𝑝𝑎 ) × OCR𝑚 < 0.4,
which is primarily caused by underestimating the 𝐷min gradient. The Darendeli01 model (plotted
in the dotted black line) predicts the 𝐷min gradient accurately but underestimates the 𝐷min (it is
shifted down relative to the data). This demonstrates that peat exhibits higher damping than clay
at small shear strains. The Rea17 model, shown in the dashed black line, fits 𝐷min data well and
is similar to the proposed model (Eq. 3).

Figure 8. Plot of present model for 𝑫𝐦𝐢𝐧 and models from literature against peat dataset

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 280

Model for the remaining parameters

With the models for 𝛾𝑟 and 𝐷min in place, we next utilize the entire MRD curve data to
develop the complete model. The hyperbolic relation for modulus reduction is given in Eq. (1).
The damping 𝐷(𝛾) varies with shear strain 𝛾 as
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

𝐺(𝛾) 0.1
𝐷(𝛾) = 𝐷min + 𝑏 ∗ 𝐷𝑀 (𝛾) ∗ ( ) (4)
𝐺max

where 𝑏 is a constant coefficient and 𝐷𝑀 (𝛾) is the Masing damping. The Masing damping is
entirely determined by shear modulus reduction 𝐺(𝛾)/𝐺max and takes the same formula as given
by Darendeli (2001). Coefficients 𝑎 (from Eq. 1) and 𝑏 are fitted using the peat dataset and are
found to be 0.71 and 0.65, respectively.
Model uncertainty
Representations of MRD model uncertainty are useful to estimate site response uncertainties.
Darendeli (2001) and Menq (2003) assumed 𝐺(𝛾)/𝐺max and 𝐷(𝛾) are normally distributed for
any given  and developed standard deviation models to represent this behavior. Because 𝛾𝑟 and
𝐷min are the key parameters in the MRD model, we approximate the MRD model dispersion by
the standard deviations of 𝛾𝑟 and 𝐷min at the present stage of model development. The
distributions are roughly log-normal, as shown by the distribution of 𝑙𝑜𝑔(𝛾𝑟 ) in Fig. 4. By using
a log-normal distribution, we avoid the possibility of negative simulated realizations in future
site response work. Since there is no appreciable trend of residuals with overburden stress and
OC based on Fig. 4 and Fig. 7, we propose constant standard deviations for 𝑙𝑜𝑔(𝛾𝑟 ) and
𝑙𝑜𝑔(𝐷min ) of 0.7 and 0.28, respectively.
MODEL PERFORMANCE
We compare the performance of the proposed model with two models from the literature
(Kea09, Rea17) based on overall model bias and standard deviation of residuals for shear
modulus and damping. The results are summarized in Table 2. Because the recommended
estimates for constant coefficients 𝑎 and 𝑏 are not optimized to minimize the overall bias but to
minimize the sum of squared errors with some constraints (to ensure the reasonable range for 𝑎
and 𝑏), the overall bias is not zero for the proposed model. But the proposed model improves the
model and has a lower bias than the other two models. Further, the standard deviation of
residuals for the proposed model is similar to Kea09 and smaller than Rea17.

Table 2. Model performance comparison among Kea09, Rea17, and this study for peat
MRD

Model Overall bias Standard deviation of residuals


This Study -0.002 0.093
Shear Modulus
Rea17 -0.006 0.106
(𝐺(𝛾𝑟 )⁄𝐺𝑚𝑎𝑥 )
Kea09 -0.041 0.093
This study 0.444 2.089
Damping
Rea17 -0.456 2.921
(𝐷(𝛾))
Kea09 0.686 2.145

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 281

SUMMARY AND CONCLUSION

We developed a new peat MRD model using a database compiled from literature. The model
was developed using a widely used framework involving hyperbolic equations for the backbone
curve that are widely used for non-peat soils and have parameters with clear physical meaning
(𝛾𝑟 and 𝑎). The two critical parameters 𝛾𝑟 and 𝐷min were also related to stress history-modified
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

effective stress using functions that can better fit data trends. The investigations showed that
organic content (OC) and peat composition (fibrous structure) over the parametric range of the
database (OC > 30%) do not appreciably influence these two parameters. The standard
deviations of the two parameters were also estimated. Compared to the previous peat MRD
models, the proposed model has more easily interpretable parameters and reduces model
dispersion.

ACKNOWLEDGMENTS

This study was supported by the California Department of Water Resources (Contract No.
4600012415). We gratefully acknowledge this support. We appreciate constructive comments on
the model from Tadahiro Kishida, Jamison Steidl, Ivan Wong, and Albert Kottke.

REFERENCES

Boulanger, R. W., Arulnathan, R., Harder, L. F., Jr., Torres, R. A., and Driller, M. W. (1998).
Dynamic properties of Sherman Island peat. J. Geotech. Geoenviron. Eng., 124 (1): 12–
20.
Darendeli, M. B. (2001). Development of a new family of normalized modulus reduction and
material damping curves. Ph.D. Thesis. Dept. of Civil Engineering, Univ. of Texas, Austin,
Texas.
Kallioglou, P., Tika, T., Koninis, G., Papadopoulos, S., and Pitilakis, K. (2008). Shear modulus
and damping ratio of organic soils. Geotech. Geol. Eng., 27 (2): 217.
Kishida, T., Boulanger, R. W., Abrahamson, N. A., Wehling, T. M., and Driller, M. W. (2009a).
Regression models for dynamic properties of highly organic soils. J. Geotech. Geoenviron.
Eng. 135 (4): 533–543.
Kishida, T., Wehling, T. M., Boulanger, R. W., Driller, M. W., and Stokoe, K. H., II. (2009b).
Dynamic properties of highly organic soils from Montezuma Slough and Clifton Court. J.
Geotech. Geoenviron. Eng. 135 (4): 525–532.
Kramer, S. L. (2000). Dynamic response of mercer slough peat. J. Geotech. Geoenviron. Eng.,
126 (6): 504–510.
Menq, F., (2003). Dynamic properties of sandy and gravelly soils. Ph.D. Thesis. Dept. of Civil
Engineering, Univ. of Texas, Austin, Texas.
Rodriguez-Marek, A., Kruiver, P. P., Meijers, P., Bommer, J. J., Dost, B., Elk, J., and Doornhof,
D. (2017). A regional site-response model for the Groningen gas field. Bull. Seismol. Soc.
Am., 107 (5): 2067–2077.
Tokimatsu, K., and Sekiguchi, T. (2006). Effects of nonlinear properties of surface soils on
strong ground motions recorded in Ojiya during 2004 mid Niigata Prefecture earthquake.
Soils Found., 46 (6): 765–775.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 282

Wehling, T. M., Boulanger, R. W., Arulnathan, R., Harder, L. F., and Driller, M. W. (2003).
Nonlinear dynamic properties of a fibrous organic soil. J. Geotech. Geoenviron. Eng. 129
(10): 929–939.
Zwanenburg, C., Konstadinou, M., Meijers, P., Goudarzy, M., König, D., Dyvik, R., Carlton, B.,
van Elk, J., Doornhof, D., and Korff, M. (2020). Assessment of the dynamic properties of
Holocene peat. J. Geotech. Geoenviron. Eng. 146 (7): 04020049.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 283

Study of Undrained Shear Strength of Laponite for Use as Transparent Clay Surrogate

Abdurrahman Almikati, S.M.ASCE1; Jorge G. Zornberg, P.E., F.ASCE2;


and Rodrigo Cesar Pierozan3
1
Postdoctoral Research Fellow, Dept. of Civil, Architectural, and Environmental Engineering,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Univ. of Texas at Austin. Email: almikati@utexas.edu


2
Priddy Centennial Professor, Dept. of Civil, Architectural, and Environmental Engineering,
Univ. of Texas at Austin. Email: zornberg@mail.utexas.edu
3
Postdoctoral Research Fellow, Dept. of Civil and Environmental Engineering, Pontifical
Catholic Univ. of Rio de Janeiro, Rio de Janeiro, Brazil. Email: rodrigopierozan@hotmail.com

ABSTRACT

The use of so-called “transparent soils” as proxy geotechnical materials has allowed for the
non-intrusive observation of a variety of models representing different engineered systems.
Laponite is one such soft clay surrogate that has seen increased usage in recent years. However,
this material has yet to be subjected to a thorough characterization and quantification of its
physical and mechanical properties. The study presented herein followed a systematic approach
towards the identification of the undrained shear strengths of laponite mixtures across a wide
range of mixes and additive dosages. Results from vane shear tests showed that the undrained
shear strength increased with (1) increasing laponite colloid content, (2) decreasing additive
dosage, and (3) aging time. The base geotechnical characteristics and mechanical properties of
the clay surrogate as provided in this study are expected to facilitate proper interpretation of the
behavior of this surrogate material in geotechnical physical models involving transparent clays.

INTRODUCTION AND BACKGROUND

Transparent soils have been the subject of a number of geotechnical studies in recent years.
Such materials have been used to visualize variations in the internal structure of soils during
loading, including the associated localized internal displacements. This can be achieved by
combining the transparent nature of these soils with the use of seeding particles and laser-based
tracking tools and techniques.
Consistent with the recent advances, transparent soil surrogates can be grouped into four
categories: (1) amorphous silica powder that behaves like clays (Iskander et al. 1994); (2)
transparent silica gels that behaves like sands (Iskander et al. 2002); (3) Aquabeads that can be
used to model flow in porous media (Tabe 2015); and (4) fused quartz, suitable for modeling the
geotechnical properties of sands (Ezzein and Bathurst 2011). More recently, a fifth group known
as clay surrogates has emerged and involves the use of transparent synthetic clays (Beemer and
Aubeny 2012), which involves the use of laponite colloids, a manufactured, synthetic product
involving magnesium lithium phyllosilicate (MLPS) that is similar to natural hectorite clays
(e.g., Wallace and Rutherford 2015, Beemer et al. 2016, Ads et al. 2020a).
When mixed with water, MLPS white powder results in a transparent synthetic clay material.
This material exhibits very low undrained shear strength values, comparable to those of soft
marine clays. A series of published studies have documented efforts to establish geotechnical
properties of MLPS mixes (Wallace and Rutherford 2015), assess their transparency (Yi et al.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 284

2018, Pierozan et al. 2021), and quantify their shear strengths using miniature ball penetrometer
tests (Ads et al. 2020a). MLPS was also used in physical laboratory-scale models of geotechnical
systems that incorporated laser techniques and image processing to visualize failure surfaces
developed upon loading caisson foundations (Wallace and Rutherford 2017).
A practical limitation facing the use of MLPS is the difficulty of preparing mixtures with
high solid concentrations, partly because of the comparatively rapid gelation process. In order to
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

address this issue, the use of additives has been adopted during sample preparation to retard the
time of gelation upon mixing the MLPS powder with water (Ruzicka and Zaccarelli 2011;
Beemer et al. 2016). Specifically, Sodium Pyrophosphate (SPP) has been used in recent studies
to facilitate the preparation of MLPS mixtures at relatively high concentrations, producing
mixtures of comparatively higher shear strength and stiffness (Ads et al. 2020b). However, only
a limited number of mixtures have been evaluated so far, so data is currently not available
regarding the strength of mixes prepared for ranges of MLPS content, SPP dosages and aging
times that are relevant for physical models in geotechnical engineering. An understanding of the
impact of SSP dosage on mix properties is particularly relevant given that this additive is key in
facilitating the preparation of mixtures at comparatively high MLPS content. In summary,
current understanding and control of the impact of increasing SPP dosage on the geotechnical
properties of MLPS, while valuable, remains limited.
This paper presents the results of a comprehensive testing program designed and executed to
better understand the parameters affecting the undrained shear strength of MLPS mixtures that
are relevant for geotechnical modeling using laponite as a clay simulant. The scope of the
experimental program includes studying the effects of MLPS content, SPP dosage and aging
time of the undrained shear strength of MLPS mixtures.

MATERIALS AND SAMPLE PREPARATION

MLPS is a synthetic layered silicate that exhibits transparency when mixed with water and
has been used in a wide range of industrial applications (Thompson and Butterworth 1992). The
specific product selected as the clay surrogate for this study is Laponite RD®, which has also
been referred in the literature as MLPS, produced by BYK Additives and Instruments (2020).
MLPS has a 2:1 layered silicate structure similar to that of the natural clay mineral Hectorite
(Van Olphen and Fripiat 1979; Pierozan et al. 2021).
In order to allow for the preparation of comparatively dense mixes (i.e., to increase the
MLPS content of the mixture), Sodium Pyrophosphate (SPP) has been used as an additive to
provide additional time for entrapped air to escape before gelation ceases (Beemer et al. 2016).
The SPP additive results in a mixture of reduced entrapped air that allows preparation of
comparatively denser MLPS mixtures. The mass percentages of SPP and MLPS used during
sample preparation are quantified as follows:
𝑚𝑆𝑃𝑃
𝐶𝑆𝑃𝑃 (%) = 𝑚 × 100 (1)
𝑤 +𝑚𝑆𝑃𝑃 +𝑚𝑙𝑎𝑝

𝑚𝑙𝑎𝑝
𝐶𝑙𝑎𝑝 (%) = 𝑚 × 100 (2)
𝑤 +𝑚𝑆𝑃𝑃 +𝑚𝑙𝑎𝑝

where CSPP represents the dosage of the rheological additive, Clap represents the MLPS content,
mSPP represents the oven-dried mass of SPP, mlap represents the oven-dried mass of MLPS, and

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 285

mw represents the mass of distilled water. The amount of water that was added to prepare the
mixtures, in order to achieve the total mas of water mw, was defined considering the initial
moisture content in SPP (68%) and MLPS (8.2%).
The MLPS mixtures in this study were prepared following the mixture and timing protocols
adopted by Pierozan et al. (2021), which are consistent with those proposed by BYK Additives
and Instruments (2020). Depending on the type of test conducted, mixtures were aged during 0
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(i.e., no aging), 1, 7, 14 or 28 days. Note that preparation of mixtures with MLPS content higher
than 11% was not feasible as the maximum soluble amount of SPP in distilled water is around
3.2% (Pierozan et al. 2021).

UNDRAINED SHEAR STRENGTH

A tailored vane shear test device was designed and constructed in this project to determine
the undrained shear strength of different MLPS mixtures. An experimental testing program was
subsequently conceived and executed to evaluate the effects of MLPS content, aging time, and
SPP dosage on the undrained shear strength of MLPS.
The vane blade adopted in this study was selected considering the range of particularly low
expected shear strength values. The blade measured 25.4 mm in height, 12.52 mm in diameter
and 0.762 mm in thickness. A stepper motor (Oriental Motors, model ARM46AC-T10) was used
to rotate the vane and shear the MLPS mixtures at a constant speed of 1 rad/min (~60º/min), as
recommended by ASTM D4648. A gauge (Mark-10, model MTT03-10Z) with a capacity of 0.07
N.m was used to control the torque during testing. A leveling table was used to lift the sample to
the required position to minimize sample disturbance. Readings were taken at a depth of 60 mm
(approximately 5 times blade diameter) below the sample surface, which is consistent with the
minimum depth of 1 blade diameter specified in ASTM D4648. A displacement of 1 radian was
selected as reference for all tests. Figure 1 depicts the vane shear setup developed in this
investigation.
The MLPS content (by dry mass) of the samples tested in this experimental program ranged
from 2% to 11%, while the additive dosage ranged from 0.00% to 3.30%. The various mixtures
and samples were prepared following the mixing protocols described by Pierozan et al (2021).
After mixing, each batch of MLPS slurry was poured into a 600 mL beaker with internal
diameter meeting the free border requirements specified in ASTM D4648 to minimize boundary
effects. The mixtures were subsequently covered with Saran wrap to avoid desiccation and were
left to age for predetermined periods of time (1, 7, 14 and 28 days).

Effect of MLPS Content

A series of laboratory vane shear tests were conducted on 30 transparent clay mixtures
involving varying MLPS contents and SPP dosages. The MLPS content ranged from 2 to 11%,
which corresponds to MLPS contents that fall in the range of previously reported studies in the
technical literature. The percentage of dry SPP used in the testing program ranged from 0 to
3.3%. Distilled water was used in all the mixtures prepared in this study. It should be noted that
the maximum soluble concentration of SPP in distilled water is 3.30%. For each mixture, four
identical samples were prepared in separate molds for testing after aging for 1, 7, 14 and 28 days
under a constant temperature of 25°𝐶.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 286
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Vane shear setup used for determination of undrained shear strength of different
MLPS mixes.

Table 1 provides a summary of the peak 𝑆𝑢 values obtained for the 120 shear strength tests
conducted as part of this study. Previously reported studies (e.g., Wallace and Rutherford 2015)
had involved tests conducted using mixtures prepared with MLPS content values of 4% and
4.5% without the addition of SPP, which ultimately resulted in a peak strength of around 0.5 kPa.
Results from shear strength tests carried out in this study and summarized in Table 1, which
indicate peak 𝑆𝑢 values often exceeding 1 kPa and reaching up to 2 kPa for mixtures with
comparatively high MLPS content achieved by using SPP additive. These comparatively high
𝑆𝑢 values were obtained after as early as seven days of aging (e.g., mixtures L9-S0.42, L11-
S0.10, L11-S0.41 and L11-S1.58). The effect of one of the relevant variables, MLPS content,
was quantified by comparing the 𝑆𝑢 results obtained at a constant temperature (i.e., 25℃), the
same aging period, and similar ranges of SPP dosage. The results of such comparison are
presented in Figure 2, which indicate that the 𝑆𝑢 shows an increasing trend with increasing
MLPS content for a given aging period and SPP dosage.
The results in Figure 2(a) correspond to the undrained shear strength 𝑆𝑢 of MLPS mixtures
prepared without SPP additive (i.e., 0% SPP content). They show a clear increase in 𝑆𝑢 with
increasing MLPS content. For example, the results for an aging period of one day indicate that
the 𝑆𝑢 approximately doubles for every 1% increase in MLPS content. Similarly increasing
trends can also be observed in this figure for aging periods of 7, 14 and 28 days.
Increasing trends in 𝑆𝑢 similar to those obtained without SPP additive can also be observed
with increasing MLPS content for the various ranges of SPP dosage considered in this study, as
observed in Figures 2(b), 2(c), 2(d) and 2(e). Also in these cases, 𝑆𝑢 approximately doubles for
every 1% percent increase in MLPS content.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 287

Table 1: Undrained shear strength for 30 mixtures over time

Su (kPa) Su (kPa)

Mix 𝐶𝑙𝑎𝑝 𝐶𝑆𝑃𝑃 1 day 7 days 14 days 28 days Mix 𝐶𝑙𝑎𝑝 𝐶𝑆𝑃𝑃 1 day 7 14 28
Label (%) (%) Label (%) (%) days days days
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

L2-S0.00 2 0.00 0.001 0.001 0.001 0.080 L6-S0.00 6 0 0.400 0.426 0.573 0.715
L3-S0.00 3 0.00 0.027 0.053 0.080 0.166 L6-S0.06 6 0.06 0.280 0.346 0.475 0.599
L3-S0.03 3 0.03 0.001 0.047 0.040 0.107 L6-S0.43 6 0.43 0.020 0.180 0.229 0.326
L3-S0.06 3 0.06 0.001 0.040 0.040 0.093 L6-S1.69 6 1.69 0.007 0.093 0.120 0.186
L3-S1.77 3 1.77 0.001 0.001 0.001 0.033 L6-S3.28 6 3.28 0.001 0.107 0.132 0.172
L4-S0.00 4 0.00 0.117 0.120 0.150 0.185 L8-S0.00 8 0 1.059 1.212 1.542 1.851
L4-S0.03 4 0.03 0.060 0.080 0.100 0.110 L8-S0.10 8 0.10 0.566 0.666 0.773 1.026
L4-S00.6 4 0.06 0.040 0.047 0.080 0.120 L8-S0.43 8 0.43 0.226 0.499 0.553 0.592
L4-S0.44 4 0.44 0.001 0.067 0.073 0.087 L8-S1.65 8 1.65 0.033 0.293 0.355 0.425
L4-S0.88 4 0.88 0.001 0.040 0.027 0.087 L8-S3.21 8 3.21 0.053 0.213 0.240 0.288
L5-S0.00 5 0.00 0.260 0.313 0.346 0.405 L9-S0.42 9 0.42 0.599 0.932 1.012 1.172
L5-S0.06 5 0.06 0.053 0.200 0.233 0.295 L11-S0.10 11 0.1 1.658 2.031 2.056 2.084
L5-S0.43 5 0.43 0.001 0.087 0.186 0.226 L11-S0.41 11 0.41 0.939 1.192 1.319 1.456
L5-S1.71 5 1.71 0.001 0.047 0.093 0.147 L11-S1.58 11 1.58 0.466 0.979 1.232 1.272
L5-S3.32 5 3.32 0.001 0.060 0.053 0.040 L11-S3.07 11 3.07 0.353 0.886 1.239 1.300

The impact on 𝑆𝑢 with increasing MLPS content is consistent with results reported by
Mourchid and Levitz (1998) regarding the effect of gelation on the strength of soft clay
surrogates. However, as shown by the results presented in Figure 2, the rate of increase of
𝑆𝑢 with MLPS content cannot be described by a single variable, as the 𝑆𝑢 depends on aspects
such as SPP dosage and aging time.
What is also important from these results is that they compare well to previous undrained
shear strength studies on MLPS samples with similar concentrations and testing conditions. Ads
et al. (2020a) conducted undrained shear strength tests on samples 4.5%, 9% and 13.5 MLPS
content, with 0% and 0.405% and 1.992% SPP dosages, respectively. The authors used
Miniature Ball Penetrometer (MBP) Tests to detect the 𝑆𝑢 variation at different aging times and
at different depths. Results from the testing campaign in this paper are comparable to the results
in Ads et al. (2020a), where the 𝑆𝑢 results for the samples L5-S0.00 and L9-S0.43 in this
research compared very well to samples with 4.5% MLPS, 0% SPP and 9% MLPS, 0.405% SPP.

Effect of SPP Dosage

In an attempt to better understand the effect of SPP dosage on the undrained shear strength of
MLPS mixtures, an evaluation was conducted by maintaining constant other variables (e.g.,
MLPS content, aging time, temperature) that also affect the 𝑆𝑢 of MLPS. The results presented in
Table 1 and Figure 2 indicate that 𝑆𝑢 decreases with increasing SPP dosage for a given MLPS
content. For example, through referring to the results in bold in Table 1, the 𝑆𝑢 reached after 28
days of aging without SPP additive is 1.85 kPa for a MLPS content of 8% [See also Figure 2(a)];
however, for the same aging period and MLPS content, the 𝑆𝑢 achieved is only 0.29 kPa when

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 288

using an SPP dosage of 3.2% [See Figure 2(e)]. The use of SPP decreases the rate of gelation,
making the sample less viscous, which in turn helps explain the decreased 𝑆𝑢 obtained with
increasing SPP dosage at any given aging period. Figure 3 shows the effect of SPP dosage on
𝑆𝑢 or mixtures prepared using different values of MLPS content and aging periods.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Undrained shear strength for increasing MLPS content after different aging
periods for SPP dosages of: (a) 0%; (b) 0.05-0.10%; (c) 0.40-0.50%; (d) 1.6-1.8%.; and (e)
3.0-3.3%.

The results in Figure 3(a) for a MLPS content of 5% and aging period of 1 day show a rapid
decrease in 𝑆𝑢 with increasing SPP dosage. For this MLPS content and aging period, the results
show that the most significant drop in shear strength occurs for comparatively low SPP dosages
(e.g., SPP< 0.5%).
Specifically, as shown in Figure 3(a), 𝑆𝑢 for Day 1 dropped about 99% for an increase in SPP
dosage from 0% to about 0.5%, with such drop reaching about 100% for an SPP dosage of 1.7%.
For aging times of 7, 14 and 28 days, the reduction in strength due to the use of SPP is about
75% for an SPP dosage of 1.7%, reaching a reduction of about 85% when the SPP dosage

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 289

increased to 3.3%. The behavior observed in Figures 3(b) and 3(c) for MLPS contents of 6 and
8% for the different aging periods is similar to that observed in Figure 3(a) for a MLPS content
of 5%.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Undrained shear strength variation with SPP dosage at different aging times for
MLPS contents of: (a) 5%; (b) 6%; (c) 8%; and (d) 11%.

While the reductions in 𝑆𝑢 with increasing SPP dosage showed similar trends for MLPS
contents below 8%, the trends were somewhat different for results obtained using a MLPS
content of 11%, as shown in Figure 3(d). As shown in this figure, the results for Day 1 show
reductions in 𝑆𝑢 of 70% for a SPP dosage increasing from 0% to 1.7%. This reduction in strength
is not as significant as the strength reduction observed for Day 1 for lower MLPS concentrations
and similar range of SPP dosage [see Figures 3(a), 3(b) and 3(c)]. Similarly, the results for Days
7, 14 and 28 show average reductions in strength of 35% and 40% for increases in SPP dosage to
1.7% and 3.3%, respectively. Also in this case, the shear strength loss obtained in mixtures with
a MLPS content of 11% is also significantly lower than the reductions (ranging from 75% to
85%) obtained when using lower MLPS contents.
It should also be noted that the rate of shear strength decrease becomes less significant with
increasing aging periods, showing the highest rate of shear strength drop at early aging times,
and lowest rate of shear strength decrease for an aging period of 28 days. This trend is further
discussed in the next section of the paper.

Effect of Aging Time

The effect of aging time on 𝑆𝑢 of MLPS mixtures can be observed in Figure 4, where the
variation of 𝑆𝑢 as function of the aging time for 10 MLPS mixtures is presented. The mixtures

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 290

shown in Figure 4 are representative of the behavior of the rest of the MLPS batches tested in
this study. These are not shown on the same graph for clarity and conciseness.

2.5

2.0
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1.5
Su (kPa)

1.0

0.5

0.0
0 5 10 15 20 25 30
Aging time (days)
L2-S0.00 L3-S0.00 L3-S0.03 L5-S0.06 L8-S0.43
L8-S1.65 L9-S0.42 L11-S0.10 L11-S0.41 L11-S1.58

Figure 4. Variation of undrained shear strength with aging time for 10 different mixtures
with varying MLPS content and SPP dosages.

The results shown in Figure 4 indicate that 𝑆𝑢 shows a clearly increasing trend with
increasing aging time. Such trend is consistent with the rheology test results presented earlier in
this paper (see Figures 2 and 3). Also, the results in Figure 4 show a decreasing rate in the time-
dependent 𝑆𝑢 gain, reaching 𝑆𝑢 value that could be considered as the ultimate value after 28 days
of aging for all mixtures evaluated in this study.
The results show that the 𝑆𝑢 at Day 7 exceeded 70% of the ultimate shear strength for most
mixtures. Accordingly, the trends in the 𝑆𝑢 obtained after testing samples aged over a seven-day
period may be considered a good indicator of the trends expected for the ultimate strength of
various MLPS mixtures.

CONCLUSIONS

This paper presents a comprehensive evaluation of the three parameters affecting the
undrained shear strength of MLPS mixtures: (1) the MLPS content; (2) the SPP dosage; and (3)
the aging time. The results of the experimental program provided significant insight on the
behavior of MLPS, which can be summarized as follows:
• The undrained shear strength of MLPS mixes was found to be governed by the MLPS
content, SPP dosage, and aging time. Specifically, the undrained shear strength was
found to increase with increasing MLPS content, decreasing SPP dosage and increasing
aging time after initial mixing.
• For the MLPS mixtures evaluated in this study, the undrained shear strength was found to
approximately double for every 1% increase in MLPS content.
• The relationship between undrained shear strength and SPP dosage relationship revealed
that, for MLPS contents below 8%, an increase in SPP dosage from 0% to 1.7% led to a

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 291

reduced, on average, the undrained shear strength by 75%; whereas an increase in SPP
dosage from 0% to 3.3% reduced the undrained shear strength by 85%. For higher MLPS
contents, the reductions in undrained shear strengths for an increase in SPP dosage from
0% to 1.7% and 3.3% were just 35% and 40%, respectively.
• The rate of undrained shear strength increase was particularly high at initial aging time.
For example, an aging time of seven days was sufficient to reach at least 70% of the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

undrained shear strengths achieved at 28 days aging time for most of the mixtures.

ACKNOWLEDGEMENTS

The authors gratefully acknowledge the funding provided by Huesker Inc. to conduct this
research.

REFERENCES

Ads, A., Iskander, M., and Bless, S. (2020a). Shear strength of a synthetic transparent soft clay
using a miniature ball penetrometer test. Geotechnical Testing Journal, 43(5), 1248-1268.
Ads, A., Iskander, M., and Bless, S. (2020b). Soil–projectile interaction during penetration of a
transparent clay simulant. Acta Geotechnica, 15(4), 815-826.
ASTM. ASTM D4648. (2013). Standard test method for laboratory miniature vane shear test for
saturated fine-grained clayey soil. West Conshohocken, PA.
Beemer, R. D., and Aubeny, C. (2012). Digital image processing of drag embedment anchors in
translucent silicate gel. Proceedings of the GeoManitoba.
Beemer, R. D., Shaughnessy, E., Ewert, K. R., Boardman, N., Biscontin, G., Aubeny, C. P., and
Grajales, F. J. (2016). The Use of Sodium Pyrophosphate to Improve a Translucent Clay
Simulate. In Geo-Chicago 2016(pp. 83-93).
BYK Additives and Instruments. (2020). Technical Information B-RI 21, Laponite –
Performance Additives. BYK Additives & Instruments, www.byk.com, Geretsried,
Germany, 24 p.
Ezzein, F. M., and Bathurst, R. J. (2011). A transparent sand for geotechnical laboratory
modeling. Geotechnical Testing Journal, 34(6), 590-601.
Iskander, M. G., Lai, J., Oswald, C. J., and Mannheimer, R. J. (1994). Development of a
transparent material to model the geotechnical properties of soils. Geotechnical Testing
Journal, 17(4), 425-433.
Iskander, M. G., Sadek, S., and Liu, J. (2002). Optical measurement of deformation using
transparent silica gel to model sand. International Journal of Physical Modelling in
Geotechnics, 2(4), 13-26.
Mongondry, P., Nicolai, T., and Tassin, J. F. (2004). Influence of pyrophosphate or polyethylene
oxide on the aggregation and gelation of aqueous laponite dispersions. Journal of colloid and
interface science, 275(1), 191-196.
Mourchid, A., and Levitz, P. (1998). Long-term gelation of laponite aqueous dispersions. The
American Physical Society, 4887-4890.
Pierozan, R. C., Almikati, A., Araujo, G. L., and Zornberg, J. G. (2021). Optical and Physical
Properties of Laponite for Use as Clay Surrogate in Geotechnical Models. ASTM
Geotechnical Testing Journal, 45(1).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 292

Ruzicka, B., and Zaccarelli, E. (2011). A fresh look at the Laponite phase diagram. Soft Matter,
7(4), 1268-1286.
Tabe, K. (2015). Transparent aquabeads to model LNAPL ganglia migration through surfactant
flushing. Geotechnical Testing Journal, 38(5), 787-804.
Thompson, D. W., and Butterworth, J. T. (1992). The nature of laponite and its aqueous
dispersions. Journal of Colloid and Interface Science, 151(1), 236-243.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Van Olphen, H., and Fripiat, J. J. (1979). Data handbook for clay materials and other non-
metallic minerals. Pergamon press, 41(42,313), 33.
Wallace, J. F., and Rutherford, C. J. (2015). Geotechnical Properties of Laponite RD. Geotech.
Test. J., 38(5), 1-14.
Wallace, J. F., and Rutherford, C. J. (2017). Response of Vertically Loaded Centrifuge Suction
Caisson Models in Soft Clay. Proc., Offshore Technology Conference 2017. 1-4 May 2017,
Houston, Texas.
Yi, L. D., Lv, H. B., Ye, T., and Zhang, Y. P. (2018). Quantification of the transparency of the
transparent soil in geotechnical modeling. Advances in Civil Engineering, 2018.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 293

Threshold Sand Content and the Behavior of Sand-Gravel Mixtures

Carmine P. Polito, Ph.D.1; Jay A. Grossman, Ph.D.2; Carter Eldridge3; Kylie Krawulski4;
William Reils5; and Grace Shebel6
1
Professor, Dept. of Civil and Environmental Engineering, Valparaiso Univ., Valparaiso, IN.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Email: carmine.polito@valpo.edu
2
Associate Professor, Dept. of Civil and Environmental Engineering, Valparaiso Univ.,
Valparaiso, IN
3
Formerly, Undergraduate Student, Dept. of Civil and Environmental Engineering, Valparaiso
Univ., Valparaiso, IN
4
Undergraduate Student, Dept. of Civil and Environmental Engineering, Valparaiso Univ.,
Valparaiso, IN
5
Formerly, Undergraduate Student, Dept. of Civil and Environmental Engineering, Valparaiso
Univ., Valparaiso, IN
6
Undergraduate Student, Dept. of Civil and Environmental Engineering, Valparaiso Univ.,
Valparaiso, IN

ABSTRACT

The threshold sand content of mixtures of sand and gravel is defined as the sand content at
which the mixture of sand and gravel transitions from behaving as a gravel to behaving as a sand.
This transition occurs when there is insufficient room in the voids created by the gravel to
contain all of the sand grains present in the mixture; at this sand content, the material begins to
transition from sand grains contained in a gravel matrix to gravel particles contained in a sand
matrix. Below the threshold sand content, the material behaves as a gravel; above the threshold
sand content, the materials behave as a sand. This phenomenon of transitional behavior has been
extensively studied for mixtures of sand and non-plastic silt, with the silt content at which the
behavioral change occurs often being referred to as the “threshold fines content.” As increasing
amounts of sand are added to a gravel, the soil mixture transitions from gravel to sandy gravel to
gravelly sand and eventually to sand. This change in soil composition leads to changes in the
maximum and minimum index void ratios (emax and emin) of the soil, as well as changes in the
coefficient of uniformity, Cu. As sand is initially added to a gravel, the range of particle sizes
increases, the index void ratios decrease and the coefficient of uniformity increases. These trends
continue until the threshold sand content is reached. Mixtures that are below the threshold sand
content typically have higher friction angles and stiffness. As the sand content continues to
increase above the threshold sand content, the range of particle sizes decreases, the index void
ratios increase and the coefficient of uniformity decreases. In this study, the changes in strength
and stiffness that are predicted to occur at the threshold sand content were investigated for four
mixtures of sand and gravel tested at different sand contents. The four mixtures studied consisted
of a uniform sand mixed with a uniform gravel, a uniform sand mixed with a poorly-graded
gravel, a well-graded sand mixed with a uniform gravel, and a well-graded sand mixed with a
poorly-graded gravel. For each combination, specimens were prepared at a constant relative
density at various sand contents and were tested in drained triaxial tests. From these tests, the
change in shear strength and stiffness that occurred as the sand content increased were
determined for each pairing of sand and gravel. At sand contents below the threshold sand

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 294

content, the strength and stiffness were similar to that of the gravel. At sand contents above the
threshold sand content, the strength and stiffness were similar to that of the sand. In each case,
the changes in friction angle, ’, and secant modulus, Es, were found to be statistically significant
and to occur at sand contents corresponding to the threshold sand content. This suggests that
there is a change in the behavior of sand-gravel mixtures that occurs at the threshold sand
content.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

INTRODUCTION

A laboratory-testing program was performed on a series of mixtures of sand and gravel, with
sand contents ranging from zero to 100%. For each soil mixture, the friction angle and
normalized secant modulus were determined. This study was performed to evaluate if there was
a statistically significant difference in behavior between the specimens with sand contents below
the lower-bound threshold sand content (LBTSC) and the specimens with sand contents above
the upper-bound threshold sand content (UBTSC).

UPPER BOUND AND LOWER BOUND THRESHOLD SAND CONTENT

As increasing amounts of sand are added to a gravel, it transitions from being a gravel to
being a sandy gravel to being a gravelly sand and eventually to being a sand. This transition
leads to a fundamental change in the soil’s behavior from gravel-like to sand-like. At similar
densities, sand-like materials have lower friction angles and lower stiffnesses than gravel-like
materials.
The transition from gravel-like behavior to sand-like behavior (and vice versa) occurs over a
relatively narrow range of sand contents. The sand content (defined as the ratio of the weight of
sand to the total weight of soil) at which this transition takes place is referred to as “the threshold
sand content”.
This type of behavioral shift has also been shown by several investigators to occur in
mixtures of sand and silt over a narrow range of silt contents (e. g. Thevanayagam 1998; Polito
and Martin 2001; Hazirbaba and Rathje 2009; Sibley and Polito 2020). This silt content is often
referred to as the “threshold fines content”.
For a mixture of sand and gravel, the gravel skeleton void ratio is the void ratio that would
exist in the soil if all of the sand grains were removed, leaving only the gravel grains to form the
soil skeleton. The gravel skeleton void ratio can range from the maximum index void ratio of the
gravel to the sand’s minimum index void ratio.
When a void in a gravel is filled with sand, the total volume of the sand (the volume of sand
grains plus the volume of the intra-sand voids) is equal to the volume of the void. The mass of
the sand in the void in the gravel varies with its density and, hence, its void ratio. For a given
void in the gravel, the mass of the sand in the void is at a minimum if the sand is at its maximum
index void ratio (i.e. its minimum density) and is at a maximum if the sand is at its minimum
index void ratio (i.e. its maximum density).
A soil with a gravel skeleton void ratio equal to its minimum index void ratio and its voids
filled with sand at the sand’s maximum index void ratio produces the smallest possible threshold
sand content. This is the lower-bound threshold sand content (LBTSC).
Conversely, a soil with a gravel skeleton void ratio equal to its maximum index void ratio
and its voids filled with sand at the sand’s minimum index void ratio produces the largest

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 295

possible sand content while still serving as a threshold fines content. This can be thought of as
the upper-bound threshold sand content (UBTSC).
Given the maximum and minimum index void ratios of the gravel fraction, it is possible to
calculate a gravel skeleton relative density corresponding to any gravel skeleton void ratio. For
any gravel skeleton relative density, there is a unique volume of voids and therefore unique
upper- and lower-bound threshold sand contents. These upper and lower bounds are a function of
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the gravel skeleton void ratio and the void ratio of the sand contained within the voids. For a
given density of sand, as the gravel skeleton relative density increases, the volume of the voids
decreases and thus both the UBTSC and the LBTSC decrease. Conversely, for a given gravel
skeleton relative density, as the density of the sand increases, the UBTSC and the LBTSC
increase. These relationships are shown in Figure 1.
For any void ratio of the gravel skeleton and any void ratio of the fines, the UBTSC and
LBTSC can be calculated using Equations 1 and 2, respectively. Figure 2 presents a plot of the
UBTSC and LBTSC as a function of gravel skeleton relative density for mixtures of uniform
gravel and Long Lost Lake sand.

Gss (eg,max )
UBTSC= (1)
Gss (eg,max )+Gsg (1+es,min )

Gss (eg,min )
LBTSC= (2)
Gss (eg,min )+Gsg (1+es,max )

Where: Gss = specific gravity of the sand; Gsg is the specific gravity of the gravel; es,min is the
minimum index void ratio of the sand; es,max is the maximum index void ratio of the sand; eg,max
is the maximum index void ratio of the gravel and eg,min is the minimum index void ratio of the
gravel.

LABORATORY TESTING PROGRAM

In this study, a program consisting of index testing and vacuum triaxial testing was
performed on mixtures created from two sands and two gravels.
The sands used in the study were a uniform sand and a well-graded sand; the gravels used in
the study consisted of a uniform gravel and a poorly-graded gravel. The uniform sand used in the
study was Ottawa #20/30 sand, a commercially-available sand from central Illinois. It is a
medium sand with rounded particles. Its grain-size distribution curve is presented in Figure 3 and
its index properties are given in Table 1.
The well-graded sand used in the study was Long Lost Lake sand (LLL) from Clearwater
County, Minnesota. It is a coarse to fine sand with sub-rounded to sub-angular particles. Its
grain-size distribution curve is presented in Figure 3 and its index properties are given in Table
1.
The uniform gravel used in the study was derived by sieving a commercially-available coarse
aggregate from northern Indiana and removing the portion retained on the 12.7 mm (½-inch)
sieve and passing the 9.5 mm (3/8-inch) sieve. It is a medium gravel with rounded to sub-angular
particles. It has a maximum particle size of 12.7 mm and a minimum grain size of 9.5 mm. Its
grain-size distribution curve is presented in Figure 3 and its index properties are given in Table
1.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 296

The poorly-graded gravel used in the study was derived by sieving a commercially-available
coarse aggregate from northern Indiana and removing the portion passing the 12.7 mm (½-inch)
sieve and retained on the 9.5 mm (3/8-inch) sieve. It is a medium to fine gravel with rounded to
sub-angular particles. It has a maximum particle size of 16 mm and a minimum particle size of
4.8 mm. Its grain-size distribution curve is presented in Figure 3 and its index properties are
given in Table 1.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1: Graphical representation of threshold sand content relationships

Figure 2: Variation of upper- and lower-bound threshold sand contents with gravel
skeleton relative density for mixtures of Long Lost Lake sand and uniform gravel

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 297

Table 1: Index Properties of the Soils Tested

Soil Gs D50 Cu emax emin ’ Es/3’


(mm)
Uniform sand 2.65 0.71 1.20 0.647 0.431 31.2o 161
LLL sand 2.63 0.55 4.71 0.577 0.300 32.4o 130
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Uniform gravel 2.50 11.0 1.16 0.671 0.433 34.7o 204


Poorly-graded 2.50 7.0 1.53 0.593 0.409 35.8o 224
gravel

The upper-bound and lower-bound threshold sand contents for each combination of sand and
gravel used in the study are presented in Table 2.

Table 2: Upper- and lower-bound threshold sand contents for the sand and gravel
combinations used in the study

Sand Gravel LBTSC UBTSC


Long Lost Lake Uniform 22.4% 35.2%
Long Lost Lake Poorly-graded 21.4% 32.4%
Uniform Uniform 21.8% 33.2%
Uniform Poorly-graded 20.8% 30.5%

Each of the four pairings of sand and gravel (LLL sand and uniform gravel, LLL sand and
poorly-graded gravel, uniform sand and uniform gravel, uniform sand and poorly-graded gravel)
were tested at sand contents of 100% (pure sand), 80%, 60%, 40%, 20% and 0% (pure gravel).
At each sand content, for each pairing of sand and gravel, maximum and minimum index void
ratios and specific gravities were determined.

Figure 3: Grain-size distribution curves for soils tested

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 298

Subsequently, at each sand content for each pairing of sand and gravel, three vacuum triaxial
tests were performed resulting in a total of 60 vacuum triaxial tests. Specimens with a diameter
of 100 mm and a height of 190 mm were prepared by dry deposition to a relative density of 40%.
The 100 mm diameter was chosen for the specimens because the maximum particle size tested
was 16 mm. As a result, the ratio of specimen diameter to particle diameter was approximately
six, which is considered the upper limit for triaxial testing. Dry deposition was used for specimen
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

preparation to minimize segregation of the soil.


The vacuum triaxial tests were performed at confining stress of 35, 60 and 85 kPa. The
triaxial tests were run to an axial strain of 10% and the maximum deviator stress was used as the
failure criterion. Mohr-Coulomb failure envelopes were plotted to determine the friction angle of
the mixture. Typical results from the tests are shown in Figure 4. These results are for the
mixture of 60% uniform sand and 40% uniform gravel.
Additionally, for each test, the stiffness of the soil was represented by its normalized secant
modulus, where secant modulus was defined as the slope of a line connecting the origin of the
stress-strain curve and the point on the curve at an axial strain of 2% (Lambe and Whitman
1969). In order to account for the effect of different confining stress, the secant modulus was
normalized by dividing the secant modulus determined in a test by the effective confining stress
used in that test. For each soil mixture, a mean normalized secant modulus was calculated by
averaging the values of normalized secant modulus from each of the three triaxial tests.
Typical results from the triaxial test program are presented in Figure 5, which shows the
variation of friction angle and normalized secant modulus with sand content for mixtures of
uniform sand and uniform gravel.

160
35 kPa 60 kPa 85 kPa
Peak Deviator Stress (kPa)

300 140
Deviator Stress (kPa)

y = 0.6377x
250
120 ' = 32.2o
100 R² = 0.9985
200
80
150
60
100 40

50 20

0 0
0 2 4 6 8 10 0 50 100 150 200 250
Axial Strain (%) Confining Stress (kPa)

Figure 4: Triaxial test results (a) and Mohr-Coulomb failure envelope (b) for the mixture
of 60% uniform sand and 40% uniform gravel

RESULTS

The results from the testing program are presented in Tables 3 and 4. These results were
divided into two groups for each pairing of sand and gravel: soil mixtures with sand contents
below the LBTSC and soil mixtures with sand contents above the UBTSC. For each range, the
average and the coefficient of variation (CoV) were calculated for the friction angle and the
normalized secant modulus.
The results were then evaluated by performing hypothesis tests using one-tailed t-tests
assuming unequal variances to compare the mean value of the specimens with sand contents
below the LBTSC and the mean value of the specimens with sand contents above the UBTSC.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 299

The null hypothesis was that the means of the two groups were the same. The alpha value for the
hypothesis testing was set at 0.05. The results of the hypothesis testing are presented in Table 5
and discussed in the Discussion of Results section.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5: Variation of friction angle and normalized secant modulus with sand content for
mixtures of uniform sand and uniform gravel

Table 3: Results Triaxial Tests: Friction Angles

Below LBTSC Above UBTSC


Sand Gravel Mean ’ CoV Mean ’ CoV
o o
Long Lost Lake Uniform 34.6 0.012 33.7 0.020
Poorly-
Long Lost Lake 35.7o 0.006 33.5o 0.039
graded
Uniform Uniform 33.4o 0.055 31.3o 0.032
Poorly-
Uniform 34.6o 0.051 30.6o 0.088
graded

DISCUSSION OF RESULTS

As can also be seen in Table 5, the null hypothesis is rejected for the friction angles of all
four pairings of sand and gravel. This means that it is highly likely that the friction angles for soil
mixtures with sand contents below the LBTSC (which exhibit gravel-like behavior) are higher
than the friction angles for soil mixtures with sand contents above the UBTSC (which exhibit
sand-like behavior).
As can also be seen in Table 5, the null hypothesis is rejected for the normalized secant
moduli of all four pairings of sand and gravel. This means that it is highly likely that the
normalized secant moduli for soil mixtures with sand contents below the LBTSC (which exhibit
gravel-like behavior) are larger than the normalized secant moduli for soil mixtures with sand
contents above the UBTSC (which exhibit sand-like behavior).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 300

Table 4: Results Triaxial Tests: Secant Modulus

Below LBTSC Above UBTSC


Mean Mean
Sand Gravel CoV CoV
Es/3’ Es/3’
Long Lost Lake Uniform 175.7 0.200 151.7 0.119
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Long Lost Lake Poorly-graded 182.7 0.321 143.4 0.152


Uniform Uniform 204.2 0.002 159.6 0.096
Uniform Poorly-graded 173.6 0.412 137.4 0.174

Table 5: Results of Hypotheses Testing

Sand Gravel Friction Angle Normalized Secant Modulus


Long Lost Lake Uniform Reject null hypothesis Reject null hypothesis
Long Lost Lake Poorly-graded Reject null hypothesis Reject null hypothesis
Uniform Uniform Reject null hypothesis Reject null hypothesis
Uniform Poorly-graded Reject null hypothesis Reject null hypothesis

CONCLUSIONS

A testing program was performed on mixtures of sand and gravel with sand contents ranging
from zero to 100% prepared by dry deposition to a relative density of 40%. For each soil
mixture, the friction angle and the normalized secant modulus were determined through vacuum
triaxial testing. Hypothesis testing was then used to determine whether there was a statistically
significant behavioral difference between soil mixtures with sand contents below the lower-
bound threshold sand content and soil mixtures with sand contents above the upper-bound
threshold sand content.
From this study, the following conclusions were drawn:
• The value of friction angle for mixtures with sand contents below the LBTSC is
independent of sand content. This is because the soil’s behavior for these sand contents is
dominated by gravel particle to gravel particle contact with most of the sand remaining in
the pore spaces formed between the gravel particles.
• The value of friction angle for mixtures with sand contents above the UBTSC is
independent of sand content. This is because the soil’s behavior for these sand contents is
dominated by sand grain to sand grain contact with the gravel particles isolated within the
sand matrix.
• The strength (as measured by the friction angle) for mixtures with sand contents above
the UBTSC was found to be lower than the value of friction angle for mixtures with sand
contents below the LBTSC. This indicates that a statistically significant behavioral
change in strength occurs as the sand content of the soil crosses the threshold sand
content.
• The stiffness, (as measured by the normalized secant modulus), for mixtures with sand
contents above the UBTSC was found to be lower than the value of friction angle for
mixtures with sand contents below the LBTSC. This indicates that a statistically
significant change in stiffness occurs as the sand content of the soil crosses the threshold
sand content.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 301

Overall, the results of this study indicate that it is highly probable that mixtures of sand and
gravel with sand contents below the lower-bound threshold sand content have different
properties of strength and stiffness than mixtures of sand and gravel with sand contents above the
upper-bound threshold sand content.

ACKNOWLEDGEMENTS
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

The first author would like to thank Valparaiso University for the financial support provided
through the Alfred W. Sieving Chair of Engineering. The authors would also like to thank Greg
Scherzer for his assistance in obtaining samples of Long Lost Lake Sand.

REFERENCES

Hazirbaba, K., and Rathje, E. M. (2009). “Pore Pressure Generation of Silty Sands due to
Induced Cyclic Shear Strains,” Journal of Geotechnical and Geoenvironmental Engineering,
135(12):1892-1905.
Lambe, T. W., and Whitman, R. V. (1969). Soil Mechanics. John Wiley & Sons, New York.
Polito, C. P., and Martin, J. R. (2001). “The Effects of Non-Plastic Fines on the Liquefaction
Resistance of Sands,” Journal of Geotechnical and Geoenvironmental Engineering 127(5),
408-415.
Polito, C., and Sibley, E. (2020). “Threshold Fines Content and the Behavior of Sands with Non-
Plastic Silts.” Canadian Geotechnical Journal 57(3), 462-465.
Thevanayagam, S. (1998). “Effect of Fines and Confining Stress on Undrained Shear Strength of
Silty Sands,” Journal of Geotechnical and Geoenvironmental Engineering 124(6), 479–491.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 302

Effect of Particle Size Distribution on Monotonic Shear Strength and Stress-Dilatancy of


Coarse-Grained Soils

Mandeep Singh Basson, S.M.ASCE1; Alejandro Martinez, Ph.D., A.M.ASCE2;


and Jason T. DeJong, Ph.D., F.ASCE3
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Dept. of Civil and Environmental Engineering, Univ. of California, Davis, CA.
Email: mbasson@ucdavis.edu
2
Dept. of Civil and Environmental Engineering, Univ. of California, Davis, CA.
Email: amart@ucdavis.edu
3
Dept. of Civil and Environmental Engineering, Univ. of California, Davis, CA.
Email: jdejong@ucdavis.edu

ABSTRACT

Natural soil deposits can consist of particles with a wide range of sizes. In current practice,
the assessment of shear strength and stress-dilatancy behavior of coarse-grained soils is based on
methods developed for poorly graded sands, without explicit consideration for differences in
gradation. This paper investigates the influence of the range of particle sizes on the monotonic
shear strength and the stress-dilatancy response of poorly- to well-graded soils. Using the 3D
discrete element method (DEM), the applicability of commonly used sand-based stress-dilatancy
frameworks is assessed for a range of gradations. This DEM investigation employs clumps of
spheres to accurately simulate the particle shapes on specimens with coefficients of uniformity
(CU) varying between 1.9 and 6.9. These specimens were subjected to isotropically consolidated
drained triaxial compression at various relative densities and confining stresses with the
objective of isolating the effects of particle size distribution from those of particle shape. The
peak and critical state shear strengths and the dilatancy responses of the specimens with different
gradations are evaluated. For the same state parameter, the results indicate an increase in the
shear strength and rate of dilation as the range of particle sizes increases. However, the critical
state line shifts downward, and its slope decreases as CU is increased. The DEM results are
compared to Bolton’s stress-dilatancy relationship to highlight the inadequacies of using clean
sand-based frameworks in capturing the behavior of well-graded soils.

INTRODUCTION

The mechanical behavior of sandy and gravelly soils is typically governed by a variety of
factors, including mineralogy, particle size distribution, particle shape, stress history, initial state,
relative density, and inherent fabric. Widely adopted stress-dilatancy theories, presented in
seminal works of Rowe (1962) and Bolton (1986), are typically based on data acquired from
poorly-graded sandy soils. Several variations of these relationships have been proposed to
capture the various salient features of granular material behavior. However, most of these
relationships are still based on experimental data obtained for poorly graded soils (Houlsby 1991,
Vaid and Sasitharan 1992, Chakraborty and Salgado 2010, Harehdasht et al. 2017). Due to a
paucity of data on the behavior of well-graded soils, partly due to sample size constraints in
experimental testing and computational constraints in numerical simulations, sand-based
engineering approaches have been adopted to predict the response of well-graded coarse-grained

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 303

soils. Overall, these prediction procedures lead to an increase in uncertainty in the response of
soils supporting critical infrastructure.
Several researchers have recently investigated the variations in stress-dilatancy and volume
change of well-graded gravelly soils with high Cu values. Although the published results of these
investigations advance our understanding of the behavior of well-graded soils, there are still
knowledge gaps and limitations. Studies have shown that increasing Cu reduces the maximum
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

and minimum void ratios, changing the relative location and slope of the critical state line in the
void ratio versus mean effective stress space (Youd 1973, Fragaszy et al. 1990, Li at al. 2014,
Ahmed 2019). The peak shear strength and dilative volumetric response increase with an
increase in Cu for drained monotonic testing (Simoni and Houslby 2006, Hamidi et al. 2012,
Harehdasht et al. 2017). However, greater Cu soils exhibit more contractive behavior, lower
undrained strengths, and higher liquefaction potential when compared at the same void ratio (Liu
et al. 2014, Li et al. 2015). Contrarily, Carey et al. (2022) report higher resistance to generation
and faster dissipation of pore pressures in high Cu soils when the relative density (DR) is held
constant. Some studies consider coarser particles to be floating in the overall soil matrix, where
variation in coarse fraction has minimal impact on the overall response (Fragaszy et al. 1992). In
contrast, monotonic DSS tests on angular sands have shown a decline in shearing strength as
coarser particle concentration increases (Chang and Phatachang 2016). From a micromechanical
standpoint, coarser particles have higher coordination numbers in the packing and can carry the
most force, implying a more significant effect on the overall behavior of well-graded soils
(Wood and Maeda 2008, Li et al. 2015, Kuei et al. 2020). The lack of agreement on the influence
of Cu on macro- and micro-scale behavior adds to the ambiguity in soil response and can lead to
over-conservative designs.
In this paper, the influence of gradation on the shear strength and stress-dilatancy behavior of
well-graded soils is investigated using the discrete element method (DEM). A series of
monotonic drained triaxial simulations were performed on specimens with poorly- to well-
graded gradations with CU ranging from 1.9 to 6.9 over a range of state parameters. DEM is
used to isolate the effect of gradation from those of particle shape and remove the effect of strain
localizations, inertial effects, specimen size, and particle crushing. The overall aim of the work is
to: (i) describe trends in the change in shear strength, stress-dilatancy, and volume change with
soil gradationand (ii) evaluate the applicability of sand-based frameworks presented in Bolton
(1986) to the response of well-graded soils.

SIMULATION METHODOLOGY

Simulated Materials

The particle shape and gradations used in this study are based on naturally occurring coarse-
grained soil sourced from a marine deposit in Mauricetown, New Jersey. The sourced soil
consisted of sub-angular quartz particles and was sieved into four poorly graded soils, namely
100A, 100B, 100C, and 100D, with A having the smallest median particle size (D50) and D
having the largest D50. These soils were mixed in different proportions to create well-graded soil
mixtures with a range of Cu values while minimizing the variability due to mineralogy and
particle shape. The particle size distributions from two such mixes, 25ABCD and 33ABC, were
used in this study, where the name denotes the names of the poorly graded soils used in the
mixture and their mass proportion. In addition, the poorly graded soil A was also used. These

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 304

soils have been extensively researched using physical (Sturm 2019, Carey et al. 2022), laboratory
(Ahmed 2021, Reardon et al. 2022), and numerical (Kuei et al. 2020, Chiaradonna et al. 2022)
modeling at UC Davis. The gradations of these soils was replicated in the DEM simulations. The
grain sizes were upscaled by 20 times to achieve realistic computational times (Kuei et al. 2020)
Figure 1(a) shows the grain size distributions of simulated materials, and Table 1 summarizes the
properties of the tested gradations.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Table 1. Average properties of the simulated gradations

D10 D30 D50 D60


Soil Cu Cc emax emin
(mm) (mm) (mm) (mm)
100A 2.1 2.6 3.4 4.1 1.95 0.85 0.94 0.62
33ABC 3.2 6.3 10.4 14.8 4.78 0.87 0.69 0.45
25ABCD 4.1 8.2 17.6 28.2 6.94 0.66 0.62 0.39

Figure 1. (a) Grain size distributions of the simulated gradations and (b) the variation of
emax and emin with an increase in Cu.

Simulation methodology

The 3D DEM code YADE was used to simulate drained monotonic triaxial tests on
specimens with particle gradations based on 100A, 25ABCD, and 33ABC. The particle shapes of
the soils were recreated in DEM using clump templates based on the particle shapes of real sand.
The real particles were scanned using a white light microscope, and the distribution of particle
shape metrics, such as sphericity and SAGI (Altuhafi and Coop, 2011), were calculated using the
MATLAB code developed by Zheng and Hryciw (2015). Two and three-particle clumps, shown
in Figure 2, were created to fit the calculated particle shape distributions. Overall, the specimen
consisted of 70% Clump 1, 10% Clump 2 and 20% Clump 3 by mass. This percentage
distribution of clumps was found by conducting a parametric study and comparing the densest
and the loosest state of the real and simulated sand at 100kPa of isotropic compression. The
variation of emax and emin with an increasing Cu from 100A to 25ABCD is presented in Figure
1(b). The plotted emax and emin were obtained by providing an inter-particle friction coefficient of

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 305

0.5 and 0.01, respectively, during the simulation. The trends of emax and emin for simulated
gradations are consistent and in range with those previously reported in experimental testing
(Youd 1973, Fragaszy 1990, Ahmed 2021).
Monotonic drained triaxial (TX) simulations were conducted on cubical specimens to
characterize the stress-dilatancy and strength of the tested gradations. Periodic boundary
conditions were employed on all the boundaries to avoid strain localization and ensure uniform
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

deformational fields within the specimens (Huang et al. 2014). A uniform strain field was
applied in the vertical direction to achieve drained conditions, whereas the confining stress was
maintained in the horizontal direction using stress-based servo control algorithms. Around 50000
clumps (120000 particles) were simulated for 100A gradation, and 75000 clumps (~190000
particles) were simulated for 25ABCD and 33ABC gradations foreach specimen. The sample
size (Dspecimen) was approximated based on the maximum the following two conditions: 20 times
D50 or 5 times the largest particle diameter (Dmax). In this study, the Dspecimen/D50 varied between
20.54 for 25ABCD and 26.11 for 100A, which are generally consistent with the
recommendations in literature (O’Sullivan 2011). A linear elastic contact model with a Mohr-
Coulomb plasticity without cohesion was used to simulate the interactions between the particles.
The normal stiffness of the contact was based on the stiffness (kn) and D50 of the particles in
contact. A particle stiffness to D50 (kn/D50) ratio of 2.5e8 N/m3, shear to normal stiffness (ks/kn)
ratio of 0.2, particle density (ρ) of 2650 kg/m3, and global damping of 0.05 were used in the
simulations. Gravity and particle crushing were not modeled in these simulations. To prepare the
specimen for TX testing, a cloud of clumped particles was isotropically compressed to a mean
effective stress of 100 kPa, 400 kPa, and 800 kPa. Once prepared, the specimen was sheared
under a constant shear rate and constant confinement to an axial strain of 25%. To achieve a
quasi-static response, the sample preparation and shearing rate were controlled such that the
inertial number was below 10-4 and the unbalanced force ratio were below 10-2 (da Cruz et al.
2005). The stresses and strains were tracked in a measurement cube with dimensions slightly
smaller than the size of the specimen. The repeatability of the simulations was checked by
rerunning specimens constructed with varying seed parameters, which influence the exact
location of particles within the specimens.

Figure 2. Clump templates created to recreate the realistic particle shape.

RESULTS AND DISCUSSION

Monotonic Drained Triaxial Response

The evolution of deviatoric stress (q) and volumetric strain (ev) with axial strain (ea) for the
three gradations, 100A, 33ABC, and 25ABCD, is illustrated in Figures 3 and 4. For brevity, only

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 306

the evolutions at tests at a confining stress of 100 kPa and 400 kPa are presented. Nonetheless,
the trends are similar for specimens tested at 800 kPa confining stress. The state parameter,
defined as the difference between the start and critical void ratios (Been and Jefferies, 1985), is
used to differentiate between the loose and dense of the critical specimens. The critical state lines
used to calculate the state parameters are shown in the following section. For both confining
stress levels, the specimens with positive state parameters show a contractive behavior without
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

mobilization of a peak shear strength, typical of loose soils. In contrast, the specimens with
negative state parameters shows a dilative behavior with a distinct peak deviatoric stresses
accompanied by strain-softening, typical of dense soils. At strains greater than 20%, the
deviatoric stress of the specimens at different initial state parameters converge to the same value
and no more volume changes occur. The higher mean effective confining stress in the
simulations at 400 kPa results in greater magnitudes of deviatoric stress, delays the onset of peak
stress to a higher axial strain, decreases the rate of dilation and decreases the final volumetric
strains for all the tested gradations. For example, 25ABCD specimen with a state parameter of -
0.101 at 100 kPa reaches a peak deviatoric stress of 413 kPa at 2.98% strain with a volumetric
strain of 6.7% at critical state, while the specimen with a state parameter of -0.083 at 400 kPa
reaches a peak q of 1420 kPa at 4.52% axial strain with a volumetric strain 5.5% at critical state.
These observations indicate that the dilative tendencies are suppressed with an increase in the
confining stress.

Figure 3. Monotonic drained TX results presented as deviatoric stress-strain and


volumetric strain response at 100kPa.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 307

A comparison of the stress ratio (q/p') and volumetric strains for the specimens at near-
identical state parameters for the three gradations is presented in Figure 5. The 25ABCD and
33ABC gradations mobilize greater q/p' values than that mobilized by the 100A specimen (Fig.
5(a)). At the peak state, 25ABCD mobilizes the highest stress ratio, followed by 33ABD and
100A, respectively. However, 100A has the highest stress ratio at the critical state, followed by
33ABCD and 25ABCD, respectively. The volumetric strains follow a similar pattern, with
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

25ABCD and 33ABC showing a stronger dilative response than 100A and a higher dilation rate
and final volumetric strains (Fig. 5(b)). The dilation rate and volumetric strains for 25ABCD are
the greatest, followed by 33ABC and 100A, indicating that the wider gradation produces a more
dilative response.

Figure 4. Monotonic drained TX results presented as deviatoric stress-strain and


volumetric strain response at 400kPa.

Figure 5. Comparison of (a) stress ratio, (b) volumetric strains for the tested gradations at
100 kPa.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 308

Critical State Lines

The critical state lines (CSL) in the q-p' and the e-log(p') plane were estimated using the
average of the measurements obtained for the last 1% strain for each triaxial test. The critical
state lines for 100A, 33ABC, and 25ABCD based on tests at confining pressures of 100kPa,
400kPa, and 800kPa are presented in Figure 6. The critical state friction angle, ϕ'cs, was
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

calculated by fitting the critical state points in the q-p' space with a straight line passing through
the origin to define the slope M. The peak angle, ϕ'p, was calculated using the maximum q/p'
mobilized in each test. A clear curvature can be observed for the CSL in the e-log(p') space, due
to which the CSL was approximated using the power function as follows:

ecs = eref – λ (p'/ pat) ξ (1)

where eref is the reference void ratio controlling the position in the CSL, λ controls the slope of
the CSL, pat is atmospheric pressure (101.325 kPa), and ξ is a material constant (Wang et al.
2002). In this study, ξ is fixed at 0.7 based on recommendations from Li and Wang (1998). The
obtained critical state measurements were fitted with equation (2) using non-linear least squares
by Levenberg-Marquardt’s fitting procedure and plotted in Figure (6).
The fitted CSL indicate that the soil gradation affects the M, eref, and λ parameters. With
increasing Cu, the M slope decreases from 1.21 (ϕ'cs = 30.2°) for 100A to 1.15 (ϕ'cs = 28.8°) for
33ABC and 1.13 (ϕ'cs = 28.3°) for 25ABCD. The ϕ'cs values are in agreement with general values
for quartz sands, and the small variation in M and ϕ'cs with an increase in Cu agrees with results
presented in the literature (e.g. Hamidi et al. 2012, Harehdasht et al. 2017, Reardon et al. 2022).
The eref decreases with an increase in Cu from 0.92 for 100A to 0.59 for 25ABCD. Similarly, the
λ decreases for an increase in Cu ranging from 0.015 for 100A to 0.006 for 25ABCD. An
increase in Cu values creates a decrease in the emax and emin (Fig. 1(b)), affecting the location of
eref. Furthermore, for high Cu soils, a significant volume of small particles filling in the voids
decreases the overall compressibility of the specimen, resulting in a decrease in the λ parameter.
Similar observations for the decrease in eref and λ with an increase in Cu are presented for various
high Cu soils in the literature (e.g., Li et al. 2014, Althuafi and Coop 2011, Ahmed 2021).

Evaluation of Stress-Dilatancy Framework

The stress-dilatancy measurements obtained for 100A, 25ABCD, and 33ABC are presented
according to Bolton's (1986) framework to test the applicability of this sand-based method over a
variety of gradations. The results based on Bolton's framework are shown in Figure 7(a), which
compares the mean effective stress, p', to the difference in peak and critical state friction angles
(ϕ'p - ϕ'cs). The empirical trendlines proposed by Bolton from the dataset of 17 poorly-graded
sands (Cu < 1.9) are included for relative densities of 25, 50, 75, and 100%. The 100% DR line
delineates a specimen prepared at the emin void ratio, the densest possible state with a highly
dilative response. The values for the poorly-graded 100A mix with highly negative state
parameters (-0.206, -0.184, -0.144) are typically within the 100% density line range. However,
the data for the 33ABC and 25ABCD specimens with state parameters less than -0.1 plot at
locations higher than the Bolton trendline for DR=100%. The difference between Bolton's
prediction and observed response becomes greater with increasing confining stress. Furthermore,
the mesurements obtianed for the 25ABCD specimen exhibit the largest ϕ'p - ϕ'cs for most

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 309

confining stress, followed by 33ABC and 100A. Based on the above observations, it is
reasonable to conclude that the wider 25ABCD and 33ABC gradations lead to a more dilative
soil response than the 100A gradation and the poorly-graded soils reported in Bolton (1986).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 6. Critical state lines and respective fittings for the q-p' and e-log(p') spaces.

Bolton (1986) also modified Rowe’s stress-dilatancy relationship, between the ϕ'p - ϕ'cs and
maximum dialtion angle (ψmax), with a scalar correction (b) given as:

ϕʹp - ϕʹcs = b∙ ψmax (2)

This correction factor (b) accounts for energy losses and provides an adjustment for the
difference in the shearing mode for the direct shear and triaxial testing, with typical values for
triaxial shearing ranging from 0.3 to 0.6 (Chakraborty and Salgado). Figure 7(b) presents the
results for 100A, 25ABCD and 33ABC specimens with a dilative response. A best fit straight
line with an intercept at origin was fitted to obtain the b parameter. The fitted results show that
the parameter b is within the range of values provided in the literature, with a minimum b value
of 0.301 for 100A, an intermediate value 0.377 for 33ABC, and a maximum b value of 0.415 for
25ABCD. This increase in b with CU indicates that the existing correlations may not propertly
account for gradation variations for soils.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 310
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 7. (a) Stress-dilatancy behavior, and (b) comparison of shearing resistance versus
maximum dilatancy of monotonic drained triaxial tests as per Bolton (1986).

CONCLUSIONS

A series of monotonic drained triaxial 3D DEM simulations were performed on specimens of


three different particle size distributions to systematically investigate the effect of widening
gradation on the shear strength, volume change, and stress-dilatancy of coarse-grained soils. An
attempt was made to mimic the realistic particle shape in simulations by employing clump
templates in the DEM simulations. Evolutions of stresses and volume changes were presented
along with estimated critical state lines in the q-p' and e-log(p') planes. Test results indicate an
increase in peak shear strength, higher rate of dilations, larger volumetric strains at critical state,
and a decrease in eref and λ parameters with widening gradation. The applicability of commonly
used sand-based stress-dilatancy framework to a range of Cu levels was evaluated. Overall, while
not unreasonable, the stress-dilatancy frameworks fail to account for the significantly dilative
response of soils with wide gradations corresponding to CU values of 4.9 and 6.9. The findings
presented here can contribute to further the understanding of well-graded granular materials.
Future studies will focus on assessing the ability of liquefaction triggering frameworks to predict
the response of soils with varying gradation.

ACKNOWLEDGEMENTS

Financial support for this project comes from the National Science Foundation (NSF) under
Grant CMMI-1916152. Any opinions, findings, conclusions, or recommendations expressed in
this paper are those of the authors and do not necessarily reflect the views of NSF. The authors
would also like to acknowledge collaboration with the experimental testing team (Rachel
Reardon, Francisco Humire, Sharif Ahmed), centrifuge modeling team (Trevor Carey, Nathan
Love, Anna Chiaradonna), and numerical simulation team (Katerina Ziotopoulou) for their
contributions to the project.

REFERENCES

Ahmed, S. S. (2021). Study on Particle Shape, Size and Gradation Effects on the Mechanical
Behavior of Coarse-Grained Soils, PhD Thesis. University of California, Davis.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 311

Altuhafi, F. N., and Coop, M. R. (2011). “Changes to particle characteristics associated with the
compression of sands.” Géotechnique, 61(6), 459-471.
Been, K., and Jefferies, M. G. (1985). “A State Parameter for Sands.” Géotechnique. 35(2), 99-
112.
Bolton, M. D. (1986). “The strength and dilatancy of sands.” Géotechnique, 36(1), 65–78.
Carey, T. J., Chiaradonna, A., Love, N. C., Wilson, D. W., Ziotopoulou, K., Martinez, A., and
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

DeJong, J. T. (2022). “Effect of soil gradation on embankment response during liquefaction:


A centrifuge testing program.” Soil Dynamics and Earthquake Engineering 157, p. 107221.
Chakraborty, T., and Salgado, R. (2010). “Dilatancy and shear strength of sand at low confining
pressures.” Journal of Geotechnical and Geoenvironmental Engineering, 527–532.
Chang, W. J., and Phantachang, T. (2016). Effects of gravel content on shear resistance of
gravelly soils. Engineering Geology, 207, 78–90.
Chiaradonna, A., Ziotopoulou, K., Carey, T. J., DeJong, J. T., and Boulanger, R. W. (2022).
“Dynamic Behavior of Uniform Clean Sands: Evaluation of Predictive Capabilities in the
Element- and the System-Level Scale.” In Geo-Congress 2022. American Society of Civil
Engineers.
Da Cruz, F., Emam, S., Prochnow, M., Roux, J., and Chevoir, F. (2005). “Reophysics of dense
granular materials: Discrete simulation of plane shear flows.” Physical Review, 72, 021309.
Fragaszy, R. J., Su, W., and Siddiqi, F. H. (1990). “Effects of oversize particles on the density of
clean granular soils,” Geotechnical Testing Journal, GTJODJ, 13(2), 106-114.
Fragaszy, R. J., Su, W., Siddiqi, F. H., and Ho, C. L. (1992). “Modeling strength of sandy
gravel,” Journal of Geotechnical Engineering Division, ASCE, 118(6), 920-935.
Hamidi, A., Azini, E., and Masoudi, B. (2012). “Impact of gradation on the shear strength-
dilation behavior of well graded sand-gravel mixtures.” Scientia Iranica, 19(3), 393–402.
Harehdasht, S. A., Karray, M., Hussien, M. N., and Chekired, M. (2017). “Influence of particle
size and gradation on the stress-dilatancy behavior of granular materials during drained
triaxial compression.” International Journal of Geomechanics, 17(9).
Houlsby, G. T. (1991). “How the dilatancy of soils affects their behaviour.” In Proceedings of
the Tenth European Conference on Soil Mech. and Foundation Engg., Vol. 4. pp. 1189–
1202.
Huang, X., Hanley, K. J., O’Sullivan, C., and Kwok, F. C. (2014). “Effect of sample size on the
response of DEM samples with a realistic grading.” Particuology. 15, 107-115.
Kuei, K. C., DeJong, J. T., and Martinez, A. (2020). “Particle Size Effects on the Strength and
Fabric of Granular Media.” In Geo-Congress 2020, American Society of Civil Engineers.
Li, X. S., and Wang, Y. (1998). Linear representation of steady-state line for sand. J Journal of
Geotechnical and Geoenviromental. Engineering 12(1215), 1215–1217.
Li, G., Liu, Y., Dano, C., and Hicher, P. (2014). “Grading-Dependent Behavior of Granular
Materials: From Discrete to Continuous Modeling.” J. Eng. Mech. 141(6), 04014172.
Liu, Y., Li, G., Yin, Z., Dano, C., Hicher, P., Xia, X., and Wang, J. (2014). “Influence of grading
on the undrained behavior of granular materials.” Comptes Rendus Mecanique. 342, 85-95.
O’Sullivan, C. (2011). Particulate Discrete Element Modelling a Geomechanics Perspective
Spon Press, London.
Reardon, R., Humire, F., Ahmed, S. S., Ziotopoulou, K., Martinez, A., and DeJong, J. T. (2022).
“Effect of Gradation on the Strength and Stress-Dilatancy of Coarse-Grained Soils: A
Comparison of Monotonic Direct Simple Shear and Triaxial Tests.” In Geo-Congress 2022,
American Society of Civil Engineers.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 312

Rowe, P. W. (1962). “The stress-dilatancy relation for static equilibrium of an assembly of


particles in contact.” Proceedings of the Royal Society of London. Series A. Mathematical
and Physical Sciences, 269(1339), 500–527.
Simoni, A., and Houlsby, G. T. (2006). “The direct shear strength and dilatancy of sand-gravel
mixtures.” Geotechnical and Geological Engineering, 24(3), 523–549.
Sturm, A. P. (2019). On the liquefaction potential of gravelly soils: characterization, triggering
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

and performance. PhD Thesis. University of California, Davis.


Vaid, Y. P., and Sasitharan, S. (1992). “The strength and dilatancy of sand.” Canadian
Geotechnical Journal, 29(3), 522–526.
Wang, Z.-L., Dafalias, Y. F., Li, X.-S., and Makdisi, F. I. (2002). “State pressure index for
modelling sand behaviour.” Journal of Geotechnical and Geoenviromental. Engg.128(6),
511-519.
Wood, D. M., and Maeda, K. (2008). “Changing grading of soil: effect on critical states”. Acta
Geotechnica. 3, 3-14.
Youd, T. L. (1973). “Factors controlling maximum and minimum densities of sands. Evaluation
of Relative Density and its Role in Geotechnical Projects Involving Cohesionless Soils.” In
STP 523. ASTM International. West Conshohocken, PA. pp. 98–112.
Zheng, J., and Hryciw, R. D. (2015). “Traditional soil particle sphericity, roundness and surface
roughness by computational geometry.” Geotechnique 65(6), 494–506 (2015).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 313

Drained Clay-Pipe Interface Resistance at Low Normal Stresses and Elevated


Temperatures

M. Basma1; R. Houhou2; S. S. Najjar, M.ASCE3; and S. Sadek, M.ASCE4


1
Graduate Student, Dept. of Civil and Environmental Engineering, American Univ. of Beirut,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Beirut, Lebanon. Email: mhb30@mail.aub.edu


2
Postdoctoral Researcher, Laboratory of Soil Mechanics, Swiss Federal Institute of Technology
in Lausanne, Lausanne, Switzerland. Email: roba.houhou@epfl.ch
3
Associate Professor, Dept. of Civil and Environmental Engineering, American Univ. of Beirut,
Beirut, Lebanon. Email: sn06@aub.edu.lb
4
Professor, Dept. of Civil and Environmental Engineering, American Univ. of Beirut, Beirut,
Lebanon. Email: salah@aub.edu.lb

ABSTRACT

Offshore pipelines that transport hydrocarbons under high pressure and high temperature are
usually thermally insulated to maintain an elevated temperature and prevent any heat loss to the
surroundings. However, the temperatures at the outer-wall of the pipes may still be elevated and
therefore may potentially affect the interface resistance between the pipeline and the seabed. This
study aims at experimentally investigating the drained shear resistance of the soil-pipe interface
at elevated temperatures (22–60℃) and low normal stresses typical of field conditions (2.45–
6.1kPa). A series of direct shear tests are thus performed using a modified-for purpose direct
shear apparatus. Low and high plasticity clays, consolidated from a slurry, were sheared against
smooth and rough interfaces to characterize the peak and residual interface shear response under
drained conditions. Results indicated that the effect of elevated temperature on the interface
resistance is highly complex and dependent on the roughness of the pipe coating, the plasticity of
the clay, and the magnitude of the applied normal stress.

INTRODUCTION

Understanding the effect of temperature changes on soil behavior is increasingly being


recognized as a key factor in many engineering designs. Several applications today involve
temperature levels that can alter the soil strength. These include geothermal foundation
engineering, radioactive waste disposal, deep geothermal reservoirs, and offshore pipeline
construction. In the context of geothermal foundation engineering, many studies have been
performed to assess the influence of temperature change on soil-structure interface behavior for
energy piles. Different structural materials were tested under relatively high normal stresses
ranging between 50 and 400 kPa and different temperatures (2-60°C). The studies were
conducted using modified temperature-controlled direct shear setups, where sand and clay were
tested against concrete, steel, and porous stones. In cases involving soil-concrete interfaces,
temperature did not show any significant change on sand-pile interfaces but clearly had an
impact in clays (Di Donna et al. 2016; Maghsoodi et al. 2019; Wang et al. 2018; Yavari et al.
2016; Yazdani et al. 2019; Xiao et al. 2014). Di Donna et al. (2016) showed that the shear
strength of a normally consolidated (NC) clay-concrete interface can be improved (20-40%) with
increasing temperature and linked this improvement to thermal consolidation of the clay. In

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 314

contrast, Yavari et al. (2016) concluded that the effect of temperature on the shear strength for
clay-concrete interfaces is insignificant in the temperature range of 5 to 40℃. Yazdani et al.
(2019) evaluated the effect of heat cycles on soil– pile interface strengths under different stress
states and loading histories. For normally consolidated (NC) clays, an increase in shear strength
was reported along with a contractive volumetric behavior. Conversely a decrease in shear
strength was observed for the case of overconsolidated clays and no change in volumetric
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

behavior. Considerable effects of temperature on both friction angle and cohesion of clay-
concrete interface were also reported by Wang et al. (2018). Interface friction angles increased
with temperature and the interface cohesion rose by ~2.9kPa as temperatures were changed from
24 to 60℃. Maghsoodi et al. (2019) represented the pile by a stainless steel plate. They reported
that the thermal contraction of kaolin during heating produces a denser soil and therefore shows
a higher shear strength. Li et al. (2019) tested red clay-porous stone interfaces.Their results
indicated that the effect of temperature on the friction angle of clay and clay–structure interfaces
was insignificant, whereas notable changes were observed in the cohesion of the clay and the
adhesion between it and structure surface depending on the normal stress level and history.
Offshore pipelines are subjected to thermal loading as they operate at high temperatures and
high pressures. Under such conditions, these pipelines experience expansions and contractions
that are normally counteracted by the axial resistance between the pipe and the seabed. After
cycles of startup and shutdown events, the pipelines become susceptible to buckling and
walking. Such phenomena must be mitigated and that requires accurate estimations of the pipe-
soil interface while mimicking in-situ conditions. So far, researchers have focused on
understanding the interface behavior at low normal stresses without considering the temperature
variations. However, it has been demonstrated that the outer surface temperature of high pressure
high temperature (HPHT) pipelines can reach 55℃ (Bai et al. 2014), which may eventually alter
the interface resistance as discussed earlier. The work presented in this paper aims at determining
the drained clay-pipe interface resistance at low normal stresses and elevated temperatures
simulating the actual in situ conditions.

EXPERIMENTAL SETUP

The typical range of normal stresses in most geotechnical applications is much larger than
that acting at the pipe-soil interface in offshore applications. As such, conventional experimental
interface/shear testing devices are not suitable for offshore pipelines applications. For the work
presented herein, a direct shear device was modified to limit its inherent friction which would
cause significant errors/effects at low confinement stresses, and to incorporate thermal loadings
at the interface. The modified device was then calibrated and tested.
Modified Setup: A modified direct shear apparatus that allows for very low normal stresses,
minimal system mechanical friction, and the control of interface temperature was used to
investigate the effects of temperature change on the soil-pipe interface properties (Figure 1a).
The measurement instruments were upgraded to guarantee high precision readings of the very
low shear forces (Omega load cell capacity of 50 N and 0.05 N accuracy) and the vertical and
horizontal displacements (linear variable differential transformer (LVDT) with 0.001 mm
accuracy). The steel shear box was replaced by a custom-fabricated Teflon box with lighter
weight and lower “sliding” resistance. The conventional loading method (lever arm and
eccentricity) was improved by adding a frictionless loading frame where dead weights are
applied directly on top of the sample to ensure zero eccentricity and load uniformity. In the

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 315

modified interface direct shear test, the bottom half of the conventional shear box was replaced
with a custom aluminum base in order to withstand elevated temperatures. This base was fitted
with 3mm diameter copper tubes connected to a heating circulator pump as shown in Figure 1b.
Hot water from the circulator passes through the copper tubes in a closed loop system in order to
raise and sustain the interface’s temperature at the desired levels. The upper shear box remained
the same, accommodating a 60mm×60mm soil sample with a thickness of 10mm.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. (a) Modified direct shear setup and (b) Copper tubes configuration in lower box

Setup Calibration: The setup was calibrated to take into account the friction of the device
induced by both mechanical and thermal loading. A dummy sample made of steel was used to
carry out the calibration tests and the same arrangement and procedure followed in real tests was
adopted during calibration. Since different interface surfaces exhibit different levels of system
friction, each structure interface was tested separately. The calibration tests were performed
twice for each structure to check the repeatability of the results. The measured values of
horizontal displacement and horizontal force (measured by load cell) were recorded in each
calibration test while heating the system to 60℃ in the absence of a soil sample. As a result,
roughly 0.14 kPa of shear resistance was attributed to the heating and friction of the testing
system in the calibration test for stainless steel (SS) interfaces, and a higher shear resistance of
0.45 kPa for the case of sandpaper. These corrections were thus applied to respectively account
for the additional stresses in the interface direct shear apparatus at elevated temperatures.
Similarly, corrections were also established for the vertical displacements measured during the
heating phase. Finally, “constant temperature” conditions at the base/interface were verified by
placing three thermocouples within the shear box. Temperature uniformity across the plate was
confirmed and was shown to remain constant once the target value was reached.

TESTING MATERIALS

Soil: The soils used in the testing program were (1) a natural clay with low plasticity (LPC)
and (2) a “synthetic/mixed” clay of high plasticity (HPC), with specific gravities of 2.63 and
2.78, respectively. The LPC is a mix of 26% sand and 74% fines. It has a liquid limit LL= 29 and
a relatively low plasticity index PI=12.5. The HPC is comprised of 100% fines with all particles

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 316

passing sieve no. 200 (0.075mm), and having a liquid limit LL= 83 and PI= 54. The low
plasticity clay (LPC) is classified as CL as per the Unified Soil Classification System USCS
(ASTM D2487) and high plasticity clay (HPC) as CH.
Interface: Two types of interface materials were used to represent offshore pipelines:
stainless steel and sandpaper, representing smooth and rough interfaces, respectively. The
roughness of each interface was measured using a profilometer. The average roughness, Ra, of
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the stainless steel is 0.284μm and that of sandpaper ~2.867μm. For the rough interface tests, the
sandpaper was glued to a base stainless steel plate using epoxy. Interface test plates 100x90mm
in size were prepared for both smooth and rough conditions. They were designed to be screwed
onto the aluminum shear box base to guarantee a constant contact area during shearing.

EXPERIMENTAL PROCEDURE

Soil samples were mixed at water content equal to their respective liquid limit. The mixed
soil was then placed in a sealed container for 24 hours prior to testing to ensure homogeneity.
The soil was then placed in the 1cm thick square box in the upper part of the modified shear box.
Then, the consolidation phase was initiated by adding weights on top of the loading frame. The
weight needed was chosen based on the selected target normal stress for each test (2.45, 4.26, 6.1
and 35kPa). Once the consolidation phase was completed, the heater circulator was turned on
and hot water was pumped at a constant flow rate from the circulator through copper tubing into
the base of the shear box. After 2 hours, the maximum temperature of 60℃ was reached and
conditions maintained for 18 hours to ensure the dissipation of any generated pore pressures. In
order to account for the evaporation of water due to heating, a “water supplier” was added to the
setup assembly (Figure 1a). In this way, the water level in the shear box remained constant
throughout the test. At this point, the box screws were loosened and a gap of around 0.7 mm was
created between the interface and the upper box. Shearing tests were then performed at a slow
rate of 0.0024 mm/min to ensure drained conditions. Based on the consolidation curve, the time
to failure to ensure drained loading conditions was computed as 𝑡𝑓 = 50. 𝑡50 (ASTM-D3080),
where: tf = total estimated elapsed time to failure in seconds and t50 is the time required for the
sample to achieve 50% consolidation. The shearing rate is chosen to be less than Rd =𝑑𝑓/𝑡𝑓 so
that insignificant excess pore pressures are generated, with Rd = shear rate in mm/s and 𝑑𝑓 is the
estimated lateral displacement at failure in mm.

EXPERIMENTAL RESULTS AND INTERPRETATION

Heating Phase: Following the completion of primary consolidation under the applied normal
stress, the interface was heated from 22℃ to 60℃ while recording the vertical displacement.
Results indicated that despite the completion of primary consolidation, upward vertical
deformations due to heating (swelling) in the order of 0.05 to 0.17mm and 0.02 to 0.13mm were
respectively observed for LPC and HPC (Figure 2a and 2c) for the SS interface. This may be
attributed to the thermal expansion of water and soil (Campanella and Mitchell 1968). The
“swelling” response observed in the low pressure range (2.45kPa and 4.26kPa) was replaced by a
contractive response when the large normal stress of 35 kPa was applied. These results indicate
that the expected contractive response that is typical of normally consolidated clay at elevated
temperature (Cekerevac & Laloui 2004) may not be valid at very low normal stresses where
other factors may affect the response. For the tests involving the rough sandpaper interface,

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 317

results on Figure 2b and 2d show a slightly contractive response (vertical deformation ~


0.05mm) due to elevated temperature under normal stresses between 2.45 kPa and 6.1 kPa in
LPC. The HPC on the other hand exhibited a response ranging from slightly expansive (stress of
2.45 kPa) to slightly contractive at the larger stresses of 4.26 kPa and 6.1 kPa.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Vertical Displacement vs. time during heating phase in (a) LPC-SS, (b) LPC-
Sandpaper, (c) HPC-SS, and (d) HPC-Sandpaper

Shearing Phase: All the results are presented in Figures 3, 4, and 5 at temperatures of 22℃
and 60℃ to highlight the impact of elevated temperatures on the interface response and the effect
of the interface roughness.
1. Effect of Temperature on the Interface Response
Low Plasticity Clay (LPC): An investigation of the stress–displacement curves (Figure 3)
leads to several observations. First, the interface shear stresses at a temperature of 22℃ were
observed to exhibit peaks at horizontal displacements in the order of 0.2 to 0.5 mm, in contrast to
the elevated temperature tests (60℃) in which the response was more ductile with peaks delayed
to displacements of about 1 mm. A softening behavior was documented post peak ultimately
stabilizing as residual conditions were reached. The difference between peak and residual shear
stresses was not affected by the temperature conditions. Second, for the LPC-steel interface tests,
the peak and residual interface shear stresses were found to increase at 60℃ compared to tests at
22℃. In contrast, the LPC-rough sandpaper interface showed a reduction in both the peak and
residual stresses at 60℃. Third, the observed volumetric change as reflected through the vertical
displacement during shear was found to be highly sensitive to the temperature conditions during

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 318

the test. For the smooth interface, the LPC exhibited a contractive volumetric tendency during
shear at ambient temperature. This tendency was clearly reduced at the elevated temperature of
60℃. These results are in agreement with the experimental results of Di Donna et al. (2016) and
Yazdani et al. (2019). These researchers concluded that heating decreases the volumetric
contraction during shearing and linked this behavior to the stiffening of the NC clays that occurs
during heating. An opposite volumetric response was observed with the rough interface with
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

clear signs of dilation during shear at ambient temperature and low normal stresses. When the
rough interface was heated to 60℃, results showed that the dilative volumetric changes became
contractive at all normal stresses, indicating a significant roughness-dependent effect of
temperature on the volumetric tendencies at the clay-solid interface during shearing. Finally, the
drained Mohr-Coulomb failure envelopes were nonlinear irrespective of temperature and
roughness conditions. The nonlinearity is evidenced in the reduction in the drained residual
secant friction angle as the normal stresses increased from 2.45 to 6.1 kPa. This nonlinearity was
noted in previous studies (Najjar et al. 2007) for clays at low stresses. A comparison between the
drained secant interface friction angles (peak and residual) at ambient and elevated temperatures
indicates that for the case of the smooth interface, heating the interface to 60℃ increased the
peak and residual friction angles at all normal stresses. Interestingly, heating the rough interface
to 60℃ decreased the peak and residual friction angles.
It should be noted that the literature includes conflicting observations with regard to the
effect of temperature on the interface friction angle. For example, Di Donna et al. (2016)
indicated that the interface friction angle reduces slightly at higher temperature which is in
contrast to results from Yazdani et al. (2019) who showed that thermal loading caused a
significant increase in the interface friction angle. The experimental results of Yavari et al.
(2016) suggested that elevated temperatures may reduce or increase interface friction angles
depending on the temperature level.
High Plasticity Clay (HPC): Results in Figure 4 lead to the following main observations.
First, the interface shear response for HPC at elevated temperature is more brittle than that at
ambient temperature with peaks in shear stress exhibited at relatively smaller horizontal
displacements. Post peak, the tests at elevated temperature exhibited a brittle response that was
characterized by significant softening up to residual conditions. Second, HPC showed a marked
decrease in peak and residual shear stresses for the smooth interface at elevated temperature
compared to 22℃. The effect of temperature reduced dramatically at the highest normal stress of
6.1 kPa. Conversely, results of the rough interface tests indicated that elevated temperature did
not have a significant effect on the peak shear stress but resulted in significant reductions in the
residual stress at large deformations, particularly for the case involving normal stresses of 6.1
kPa. Third, the observed volumetric change at the interface during shearing indicated that
significant contractive and dilative vertical deformations were recorded for both stainless steel
and sandpaper at 22℃ and 60℃ respectively. Elevated temperatures led to a highly dilative
response that is indicative/typical of “overconsolidated” clay behavior. Fourth, the drained
interface Mohr-Coulomb failure envelopes for HPC were found to be slightly nonlinear for tests
conducted at ambient temperature. However, the failure envelopes at elevated temperature
exhibited a complex behavior that was sensitive to the interface roughness, normal stress, and
magnitude of shear deformations (peak versus residual). For the smooth interface, reductions
were observed in the peak and residual friction angles due to elevated temperature. For the rough
interface, heating to 60℃ increased the peak friction angles and reduced the residual friction
angles.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 319
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Normal Stress (kPa)

Normal Stress (kPa)

Figure 3. Interface response during direct shear testing at ambient temperature (22℃) and
elevated temperature (60℃) for (a) LPC with Steel and (b) LPC with Sandpaper

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 320
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Interface response during direct shear testing at ambient temperature (22℃) and
elevated temperature (60℃) (a) HPC with Steel and (b) HPC with Sandpaper

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 321

2. Effect of Pipe Roughness on the Mode of Failure


In general, the shear strength of a soil–structure interface is lower than or at most equal to the
shear resistance of the soil. As such, many researchers concluded that increasing the surface
roughness will lead to an increase in interface shear strength and the behavior becomes closer to
that of the soil (Houhou et al. 2022). These observations were confirmed in the work presented
herein even for tests that are conducted at elevated temperature. Figure 5 shows the failure
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

mechanism that was observed at the interface for LPC and HPC at elevated temperature for the
smooth and rough interfaces. Results indicate that a clean interface shear surface is observed
when LPC and HPC were sheared against the smooth stainless-steel interface with the mode of
failure being controlled by slippage between steel and clay. On the other hand, observations of
the mode of failure for the rough interface at elevated temperature indicated that clay particles
adhered to the sandpaper during the shearing phase for both LPC and HPC, indicating a
combined failure mechanism at the clay-interface.
(a) (b) (c) (d)

Figure 5. Failure mechanism of LPC on (a) Stainless steel and (b) Sandpaper and HPC on
(c) Stainless steel and (d) Sandpaper

CONCLUSION

The pipe-soil interface resistance is a key parameter in the design of HPHT pipelines. This
study aimed at investigating the impact of elevated temperature on the interface response
between clays and the pipeline coating. Tests were conducted at ambient and elevated
temperatures (22℃ and 60℃) to achieve the goals of the study. The following conclusions can be
drawn from a total of 24 interface direct shear tests that were conducted on low and high
plasticity clays tested against smooth stainless steel and rough sandpaper interfaces at normal
stresses between 2.45 and 6.1 kPa:
1. The interface direct shear response is a complex phenomenon when tested in the low
pressure range and is highly sensitive to the magnitude of the normal stress, clay
plasticity, interface roughness, and magnitude of the applied displacement.
2. Failure envelopes are nonlinear in the low pressure range.
3. The residual friction angle is found to be sensitive to increases in temperature at
relatively low normal stresses, with the sensitivity to temperature decreasing as the
normal stress is increased. This observation is important given that the residual strength
typically governs the design of pipelines in the offshore environment.
4. For highly plastic soils, where swelling occurs during heating and shearing irrespective of
the roughness, the residual resistance is lower at higher temperatures compared to
ambient temperatures. However, for low plasticity soils with lower fines content, the
behaviour is more dependent on the roughness.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 322

ACKNOWLEDGMENTS

The authors would like to acknowledge the University Research Board at the American
University of Beirut for funding this research.

REFERENCES
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Bai, Y., Niedzwecki, J. M., and Sanchez, M. (2014). Numerical investigation of thermal fields
around subsea buried pipelines. Proceedings of the International Conference on Offshore
Mechanics and Arctic Engineering - OMAE, 6B(July 2015).
Campanella, R. G., and Mitchell, J. K. Influence of temperature variations on soil behavior.
Journal of the Soil Mechanics and Foundation Division ASCE 1968; 94(SM3):709–734.
Cekerevac, C., and Laloui, L. (2004). Experimental study of thermal effects on the mechanical
behaviour of a clay. International Journal for Numerical and Analytical Methods in
Geomechanics, 28(3), 209–228.
Di Donna, A., Ferrari, A., and Laloui, L. (2016). Experimental investigations of the soil–concrete
interface: Physical mechanisms, cyclic mobilization, and behaviour at different temperatures.
Canadian Geotechnical Journal, 53(4), 659–672.
Houhou, R., Mjahed, R. B., Sadek, S., and Najjar, S. (2022). Drained Axial Pipe-Soil Resistance
at Low Confinement Using Tilt Table and Direct Shear Tests.
Li, C., Kong, G., Liu, H., and Abuel-Naga, H. (2019). Effect of temperature on behaviour of red
clay–structure interface. Canadian Geotechnical Journal, 56(1), 126–134.
Maghsoodi, S., Cuisinier, O., and Masrouri, F. (2019). Thermo-mechanical behaviour of clay-
structure interface. E3S Web of Conferences, 92, 1–6.
Najjar, S. S., Gilbert, R. B., Liedtke, E., McCarron, B., and Young, A. G. (2007). Residual Shear
Strength for Interfaces between Pipelines and Clays at Low Effective Normal Stresses.
Journal of Geotechnical and Geoenvironmental Engineering, 133(6), 695–706.
Wang, D., Lu, L., and Cui, P. (2018). Simulation of thermo-mechanical performance of pile
geothermal heat exchanger (PGHE) considering temperature-depend interface behavior.
Applied Thermal Engineering, 139(July 2017), 356–366.
Xiao, S., Suleiman, M. T., and McCartney, J. S. (2014). Shear Behavior of Silty Soil and Soil-
Structure Interface under Temperature Effects.
Yavari, N., Tang, A. M., Pereira, J. M., and Hassen, G. (2016). Effect of temperature on the
shear strength of soils and the soil–structure interface. Canadian Geotechnical Journal,
53(7), 1186–1194.
Yazdani, S., Helwany, S., and Olgun, G. (2019). Influence of temperature on soil–pile interface
shear strength. Geomechanics for Energy and the Environment, 18, 69–78.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 323

Shear and Elastic Moduli of Fine-Grained Soils: Impact of Consolidation Pressure and
Plasticity Characteristics

Beena Ajmera, Ph.D., P.E.1; Binod Tiwari, Ph.D., P.E.2; and Quoc-Hung (Bob) Phan3
1
Assistant Professor, Dept. of Civil, Construction, and Environmental Engineering, Iowa State
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Univ., Ames, IA. Email: bajmera@iastate.edu


2
Associate Vice President of Research and Sponsored Projects and Professor of Civil
Engineering, California State Univ., Fullerton, CA. Email: btiwari@fullerton.edu
3
Former Graduate Student, Dept. of Civil and Environmental Engineering, California State
Univ., Fullerton, CA

ABSTRACT

Montmorillonite, kaolinite, and quartz were used to prepare 20 samples in the laboratory to
have liquid limits ranging from 8 to 486 and plasticity indices ranging from 4 to 431. The
samples were normally consolidated to pressures ranging from 25 kPa to 100 kPa and then
subjected to bender element tests to determine the maximum Young’s modulus of elasticity and
maximum shear modulus. Both the maximum Young’s modulus of elasticity and the maximum
shear modulus were observed to depend on the consolidation pressure, the clay mineralogy, and
the plasticity characteristics. In particular, an increase in the consolidation pressure would result
in an increase in the modulus of elasticity and the shear modulus. At a constant plasticity index, a
soil containing montmorillonite had a higher moduli value than a soil containing kaolinite. An
increase in the plasticity index corresponded to a decrease in the value of the maximum Young’s
modulus of elasticity and the maximum shear modulus. Little variations were observed in the
values of the maximum Young’s modulus of elasticity and the maximum shear modulus when
the plasticity index exceeded 250 regardless of the consolidation pressure or the clay mineralogy.
Figures that can be used to estimate the maximum Young’s modulus of elasticity and the
maximum shear modulus as a function of the clay mineralogy, plasticity characteristics, and
consolidation pressure are provided.

INTRODUCTION

The maximum Young’s modulus of elasticity and the maximum shear modulus are required
parameters in a number of geotechnical engineering applications including in any numerical
modeling efforts. As a result there have been a number of studies that have proposed
relationships between various soil properties and these moduli values. Some examples include
the relationships proposed by: Hardin and Black (1966, 1968, 1969), Hardin and Richart (1963),
Iwasaki and Tatsuoko (1977), Iwasaki et al. (1978), Kokusho (1980), Kokusho et al. (1982),
Marcuson and Wahls (1972), Seed and Idriss (1970), Vardanega and Bolton (2013), Yamada et
al. (2008), and Zen et al. (1978).
Most of the studies regarding the maximum Young’s modulus of elasticity and the maximum
shear modulus available in the literature are primarily focused on the impact of one factor on the
moduli. Alternatively, the published work will look at the impact of several factors, but on a
limited number/type of soil samples. Studies that examine the influence of several factors over a
soils with a wide range of properties are lacking, which prevents researchers from obtaining a

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 324

comprehensive understanding of the maximum Young’s modulus of elasticity and the maximum
shear modulus.
This paper presents the results of bender element tests conducted on a wide range of soils
prepared in the laboratory. During these tests, both compression and shear waves were induced
in the samples to obtain measurements of the maximum Young’s modulus of elasticity and the
maximum shear modulus. The impact of clay mineralogy, plasticity characteristics and
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

consolidation pressure are evaluated. Relationships between these parameters and the maximum
Young’s modulus of elasticity or maximum shear modulus are also developed.

METHODOLOGY

In this study, twenty different soils were subjected to bender element testing. These samples
were prepared in the laboratory by mixing dry, powdered clay minerals with quartz. Specifically,
ten of the samples were mixtures of kaolinite with quartz, while the remaining ten were mixtures
of montmorillonite with quartz, as depicted in Table 1. Kaolinite and montmorillonite were
selected in this study as they are the strongest and weakest clay minerals, respectively. Several
past studies (for example, Ajmera et al. 2019, and Tiwari and Ajmera 2011a, 2011b, 2014) have
shown that the use of laboratory prepared mixtures of kaolinite, montmorillonite and quartz can
capture the static and dynamic behavior of natural soils.
Sufficient distilled, de-aired water was added to the mixture to have an initial moisture
content equal to the liquid limit. Liquid limits for each of the mixtures were measured using the
procedures in ASTM D4318. Table 1 summarizes the corresponding results. The slurry sample
was then placed in an airtight container, where it was allowed to hydrate for a period of 24 hours
before the consolidation process commenced.
Following the 24 hour hydration period, the sample was carefully placed into a custom mold.
This mold was constructed from SCH-80 PVC coupling to have an internal diameter of 73 mm
and a height of 42 mm. A schematic and additional details regarding this custom mold are
provided in Ajmera et al. (2020). While placing the slurry, attention was paid to ensure that no
air pockets were inadvertently created. The mold was placed into a water bath and sufficient
dead weights were placed on the sample to achieve an initial vertical stress of 25 kPa. The
sample was allowed to consolidate under this pressure. During this time, the vertical deformation
of the sample was recorded using a mounted caliber to allow for the development of a logarithm
of time versus displacement curve. In particular, at the start of the consolidation process,
readings were taken on more frequent intervals than after the consolidation process had
progressed for some time. The resulting curve was monitored to establish the end of the primary
consolidation phase. This process was used for all of the samples tested in this study regardless
of their plasticity index. It is noted that the samples with higher plasticity indices required more
time for the primary consolidation to be completed.
Once the primary consolidation was completed, bender elements were embedded into the top
and bottom of the sample. It is noted the custom mold was designed to allow for bender elements
to be embedded into the sample without having to remove the sample from the mold. Therefore,
minimum sample disturbance occurred. Using a WaVeMe signal generator, manufactured by
GeoComp, Inc., several different pulses were applied to the transmitter element, located at the
base of the sample. The signal received by the receiver element, at the top of the sample, was
recorded. The signal generator used in this study could induce both compression and shear waves
in the sample. At least six different compression and shear waves with varying frequencies were

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 325

applied on each sample. The minimum frequency applied in this study was 3 kHz. Some
examples of the transmitted and received wave signals are shown in Figure 1. The results in
Figure 1 are from three samples: (1) 30% kaolinite with 70% quartz, (2) 70% kaolinite with 30%
quartz, and (3) 20% montmorillonite with 80%.

Table 1. Plasticity characteristics of the twenty mixtures tested in this study.


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Kaolinite (K) – Montmorillonite (M) –


Quartz (Q) Mixtures Quartz (Q) Mixtures
K Q LL PI M Q LL PI
10 90 8 4 10 90 45 11
20 80 19 6 20 80 88 59
30 70 24 7 30 70 134 98
40 60 28 8 40 60 122 14
50 50 34 9 50 50 209 160
60 40 41 14 60 40 264 234
70 30 50 19 70 30 304 262
80 20 55 20 80 20 330 280
90 10 61 21 90 10 423 355
100 0 73 28 100 0 486 431

The time difference between the peak of the transmitted wave and the peak of the received
wave was determined. Using this and the distance between the tips of the receiver and transmitter
elements, the shear wave velocity could be calculated when a shear wave was induced, while the
compression wave velocity was calculated for each compressive wave induced in the sample.
Similar calculations were also performed using the time difference between the troughs of the
transmitted and received wave forms. Thus, twelve different measurements of both the
compression wave velocity and the shear wave velocity were obtained.
Using the values of the compression and shear wave velocities, the maximum Young’s
modulus of elasticity and the maximum shear modulus could be computed. The maximum shear
modulus is a measure of the elastic shear stiffness of the soil at small strains, while the maximum
Yong’s modulus of elasticity is a measure of the stiffness of the soil subjected to compressive
axial loads at small strains. These parameters were calculated using Equations 1 and 2 (Kramer
1996), in which Emax is the maximum Young’s modulus of elasticity, Gmax is the maximum shear
modulus, ρ is the density of the sample, vp is the compression wave velocity and vs is the shear
wave velocity. The density of the sample was obtained by measuring the weight and height of
the sample at the start of each sequence of bender element tests.

𝐸𝑚𝑎𝑥 = 𝜌𝑣𝑝2 (1)

𝐺𝑚𝑎𝑥 = 𝜌𝑣𝑠2 (2)

Due to the non-destructive nature of bender element testing combined with the minimal
disturbance that occurred during the insertion of the bender elements, the same sample was used
to obtain measurements of the maximum Young’s modulus of elasticity and the maximum shear
modulus at different consolidation pressures. Specifically, once the bender element tests were

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 326

concluded for a sample at a consolidation pressure of 25 kPa, the bender elements were removed
from the sample and additional dead weights were applied to the sample to achieve a
consolidation pressure of 50 kPa. As before, the displacement was monitored as a function of the
logarithm of time. Upon the completion of the primary consolidation phase, the sample was then
subjected to bender elements, as described above before the process was repeated with a
consolidation pressure of 100 kPa.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Example transmitted and received waves from bender element tests. Images on
the left correspond to the compression waves and those on the right are shear waves for
samples of (top) 30% kaolinite with 70% quartz, (middle) 70% kaolinite with 30% quartz,
and (bottom) 20% montmorillonite with 80% quartz. Consolidation pressure of 100 kPa.

The focus of this paper is on the maximum Young’s modulus of elasticity and maximum
shear modulus. Therefore, the compression and shear wave velocities obtained for the samples
are not discussed here. Interested readers are directed to Ajmera et al. (2020) for additional
information about these velocities. For the remainder of this paper, the average values obtained
from the twelve measurements of the maximum Young’s modulus of elasticity and maximum
shear modulus will be presented and discussed. Please note that in computing these averages, any
outliers (typically one or two measurements) in the results were discarded.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 327

RESULTS

Maximum Young’s Modulus of Elasticity. Figure 2 presents the variation of the maximum
Young’s modulus of elasticity with the plasticity index. As Figure 2 illustrates, the behavior of
the maximum Young’s modulus is found to depend on the clay mineralogy, plasticity index as
well as the consolidation pressure. Specifically, an increase in the plasticity index corresponded
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

to a decrease in the value of the maximum Young’s modulus of elasticity. However, at a given
plasticity index, the montmorillonite-quartz mixtures tended to have higher values of the
maximum Young’s modulus of elasticity than the kaolinite-quartz mixtures. Furthermore, for
any soil, an increase in the consolidation pressure resulted in an increase in the maximum
Young’s modulus of elasticity. Clay mineralogy and consolidation pressure appeared to have
little influence on the value of the maximum Young’s modulus of elasticity when the plasticity
index was greater than 250.

Figure 2. Maximum Young’s modulus of elasticity as a function of clay mineralogy,


consolidation pressure and plasticity index. M-Q and K-Q mixtures represent the mixtures
of montmorillonite with quartz and kaolinite with quartz, respectively.

Maximum Shear Modulus. The observed variation in the maximum shear modulus with
plasticity index, consolidation pressure and clay mineralogy is presented in Figure 3. The shear
moduli of the samples exhibit trends similar to those seen in the maximum Young’s modulus of
elasticity. Specifically, the maximum shear modulus decreases with an increase in the plasticity
index until a plasticity index of approximately 250 beyond which the maximum shear modulus
appears to remain constant. This reduction is more significant in the mixtures of kaolinite with
quartz than in the mixtures of montmorillonite with quartz. An increase in the consolidation
pressure also corresponds to an increase in the maximum shear modulus.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 328
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Maximum shear modulus as a function of the clay mineralogy, consolidation


pressure and plasticity index. M-Q and K-Q mixtures represent the mixtures of
montmorillonite with quartz and kaolinite with quartz, respectively.

DISCUSSION

The variations in the maximum shear modulus and maximum Young’s modulus of elasticity
may be attributed to several different factors. Density of the soil mass likely contributes to the
increases observed in the moduli values with an increase in the consolidation pressure.
Specifically, as the pore water pressure generated due to an increase in the consolidation pressure
is allowed to dissipate, the void ratio of the soil mass will decrease and the soil particles will
achieve a denser configuration. The tendency of the soil mass to deform under this denser
configuration will reduce. This, in turn, will result in an increase in the moduli of the sample.
Compressibility of the soil fabric may play in a role in causing a reduction in both the
maximum Young’s modulus of elasticity and the maximum shear modulus with an increase in
the plasticity index. Mitchell and Soga (2005) stated that the thin, flexible, platy shape of the
clay particles results in large specific surface areas. This combined with the inter-particle forces
present in clay minerals results in a compressible fabric. Several past studies include Bowels
(1979), Terzaghi and Peck (1948), Skempton (1944), Nagaraj and Srinivasa Murthy (1983) and
Tiwari and Ajmera (2011a, 2011b) have shown that the compressibility of the soil mass
increases with an increase in the plasticity index. With an increase in the compressibility of the
soil fabric, the stiffness of the soil will decrease corresponding to reductions in the maximum
Young’s modulus of elasticity and the maximum shear modulus.
Mixtures of montmorillonite with quartz were generally found to have higher values of both
the maximum Young’s modulus of elasticity and the maximum shear modulus in comparison to

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 329

mixtures of kaolinite with quartz at a constant plasticity index. These differences may be related
to the nature and type of bonds of the clay minerals. Specifically, the large electrical and
chemical bonds and repulsion forces present in montmorillonite (Mitchell and Soga 2005) may
allow this mineral to better resist the deformations applied than kaolinite. As such, it will have
higher moduli values.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

CONCLUSIONS

In this study, twenty samples were prepared in the laboratory as mixtures of montmorillonite
with quartz or mixtures of kaolinite with quartz. The samples had controlled mineralogical
compositions and a wide range of plasticity characteristics. They were subjected a series of
bender element tests to measure their maximum Young’s modulus of elasticity and maximum
shear modulus at three different consolidation pressures. The results indicated that both the
maximum Young’s modulus of elasticity and the maximum shear modulus of the samples would
decrease with an increase in the plasticity index. Furthermore, clay mineralogy and consolidation
pressure impacted the moduli values when the plasticity index was less than 250. Specifically, an
increase in the consolidation pressure would result in a denser configuration resulting in higher
values of the maximum Young’s modulus of elasticity and the maximum shear modulus. At a
constant plasticity index and consolidation pressure, the mixtures of montmorillonite with quartz
were found to have higher moduli values than the mixtures of kaolinite with quartz. These
differences may be attributed to the specific surface area and inter-particle bonding of the soil
mass.

ACKNOWLEDGEMENTS

The financial supported provided by the California State University, Fullerton Intramural
Grant is appreciated by the authors.

REFERENCES

Ajmera, B., Brandon, T., and Tiwari, B. (2019). “Characterization of the Reduction in Undrained
Shear Strength in Fine-Grained Soils due to Cyclic Loading,” Journal of Geotechnical and
Geoenvironmental Engineering, 145 (5), 04019017 1-10.
Ajmera, B., Tiwari, B., and Phan, Q.-H. (2020). “Small Strain Dynamic Properties of Silt-Clay
Mixtures,” Proceedings of Geo-Congress 2020 Geotechnical Special Publication, 318, 190-
197.
ASTM. ASTM D4318. (2010). Standard Test Methods for Liquid Limit, Plastic Limit and
Plasticity Index of Soils, ASTM International.
Bowles, J. W. (1979). Physical and Geotechnical Properties of Soils, McGraw Hill.
Hardin, B. O., and Black, W. L. (1966). “Sand Stiffness Under Various Triaxial Stresses,”
Journal of the Soil Mechanics and Foundations Division, 95 (6), 1531-1537.
Hardin, B. O., and Black, W. L. (1969). “Closure to Vibration Modulus of Normally
Consolidated Clay,” Journal of the Soil Mechanics and Foundations Division, 95 (6), 1531-
1537.
Hardin, B. O., and Richart, F. E., Jr. (1963). “Elastic Wave Velocities in Granular Soils,”
Journal of the Soil Mechanics and Foundations Division, 89 (1), 33-65.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 330

Hardin, B. O., and Black, W. L. (1968). “Vibration Modulus of Normally Consolidated Clay,”
Journal of the Soil Mechanics and Foundations Division, 94 (2), 353-369.
Iwasaki, T., and Tatsuoka, F. (1977). “Effects of Grain Size and Grading on Dynamic Shear
Moduli of Sands,” Soils and Foundations, 17 (3), 19-35.
Iwasaki, T., Tatsuoka, F., and Takagi, Y. (1978). “Shear Moduli of Sands under Cyclic Torsional
Shear Loading,” Soils and Foundations, 18 (1), 39-59.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Kokousho, T. (1980). “Cyclic Triaxial Test of Dynamic Soil Properties for Wide Strain Range,”
Soils and Foundations, 20 (2), 45-60.
Kokouso, T., Yoshida, T., and Esashi, Y. (1982). “Dynamic Properties of Soft Clay for Wide
Strain Range,” Soils and Foundations, 22 (4), 1-18.
Kramer, S. L. (1996). Geotechnical Earthquake Engineering, Prentice Hall, Upper Saddle River,
New Jersey.
Marcuson, W. E., and Wahls, H. E. (1972). “Time Effects on Dynamic Shear Modulus of Clays,”
Journal of the Soil Mechanics and Foundations Division, 98 (12), 1359-1373.
Mitchell, J. K., and Soga, K. (2005). Fundamentals of Soil Behavior, John Wiley & Sons.
Nagaraj, T. S., and Srinivasa Murthy, B. R. (1983). “Rationalization of Skempton’s
Compressibility Equation,” Géotechnique, 33 (4), 433-443.
Seed, H. B., and Idriss, I. M. (1970). “Soils Moduli and Damping Factors for Dynamic Response
Analysis,” Report No. EERC70-10, Earthquake Engineering Research Center, University of
California, Davis.
Skempton, A. W. (1944). “Notes on the Compressibility of Clays,” Quarterly Journal of the
Geological Society of London, 100, 119-135.
Terzaghi, K., and Peck, R. B. (1948). Soil Mechanics in Engineering Practice, Wiley.
Tiwari, B., and Ajmera, B. (2011a). “Consolidation and Swelling Behavior of Major Clay
Minerals and Their Mixtures,” Applied Clay Science, 54 (3-4), 264-273.
Tiwari, B., and Ajmera, B. (2011b). “New Correlation Equations for Compression Index of
Remolded Clays,” Journal of Geotechnical and Geoenvironmental Engineering, 138 (6),
757-762.
Tiwari, B., and Ajmera, B. (2014). “Effects of Saline Fluid on Compressibility of Clay
Minerals”, Environmental Geotechnics, 1 (EG2), 108-120.
Vardanega, P. J., and Bolton, M. D. (2013). “Stiffness of Clays and Silts: Normalizing Shear
Modulus and Shear Strain,” Journal of Geotechnical and Geoenvironmental Engineering,
139 (9), 1575-1589.
Yamada, S., Hyodo, M., Orense, R. P., and Dinesh, S. V. (2008). “Initial Shear Modulus of
Remolded Sand-Clay Mixtures,” Journal of Geotechnical and Geoenvironmental
Engineering, 134 (7), 960-971.
Zen, K., Umehara, Y., and Hamada, K. (1978). “Laboratory Tests and In-Situ Seismic Survey on
Vibratory Shear Modulus of Clayey Soils with Various Plasticities,” Proceedings of the 5th
Japanese Symposium on Earthquake Engineering, 721-728.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 331

A Study of Consolidation Tests on Dredged Soils with a Large Moisture Content in Coastal
Louisiana Using a Modified Odometer

Omar Shahrear Apu1 and Jay X. Wang, Ph.D., P.E.2


1
Ph.D. Student, Program of Civil Engineering, Louisiana Tech Univ., Ruston, LA.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Email: osa005@latech.edu
2
Professor, Programs of Civil Engineering and Construction Engineering Technology, Louisiana
Tech Univ., Ruston, LA. Email: xwang@latech.edu

ABSTRACT

This paper reports a comparative study of the consolidation behavior of dredged soils by
using a standard conventional odometer and a modified conventional odometer. A series of tests
were carried out on the dredged soil samples from the Louisiana coastal area. Dredged soils are
soft and highly moisturized, and thus form a large strain consolidation settlement. For this
reason, the seating load, which is the initially applied load at the consolidation test, is an
essential factor for the consolidation test of slurry-type dredged soils. The conventional odometer
comes up with a large seating pressure of 0.01 TSF (1 TSF = 107.25 kPa), which would squeeze
part of the soft slurry sample out of the odometer cell if the initial moisture content is greater
than 70%. Utilizing 3D printing technology, the dial cap was printed out using material
polylactic acid (PLA). Then, the seating pressure has been significantly reduced, and the newly
achieved seating load turned out to be 0.002 TSF, including the porous stone weight. With the
3D-printed dial cap, dredged soil samples with a moisture content up to 100% could be
successfully tested. The laboratory test results of the consolidation properties of Louisiana
dredged soils are presented in this study. The results suggest that the modified consolidation
apparatus can be used to obtain a reliable relationship between void ratio and effective stress for
materials with high initial void ratios.

INTRODUCTION

Very soft clayey sediments are likely to be produced due to the restoration process of
marshland along with coastal Louisiana. Concerning both production and utilization of such
artificial marsh, it is essential to predict the consolidation settlement, including the effect of self-
weight of dredged soils. The method of settlement analysis for such very soft soils was proposed
by (Mikasa 1963) and gradually developed by many other researchers. Umehara and Zen (1975)
and Imai (1978) previously proposed the test method based on the constant rate of strain
consolidation test. Umehara & Zen (1980) further improved the technique and testing device for
determining the consolidation constants of very soft soils and a new method to interpret test
results. Umehara & Zen (1980) introduced stainless steel and acryl consolidation ring to modify
consolidation devices for testing very soft soils. However, any standard laboratory testing
procedure for such very soft clays has not been well established yet. The conventional oedometer
test based on Terzaghi's one-dimensional consolidation theory may not be applicable to very
soft, high-water content soils due to theoretical and testing limitations. Firstly, it is very
challenging to prevent partial slurry sample from being squeezed out along the inner surface of
the consolidation ring, especially at the application of the seating load combining with the weight

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 332

of the dial cap and porous stone, which may tend to cause a significant discrepancy in the
accuracy of the test results. Secondly, large strain would happen when soil sample is
consolidated from an extremely soft state, the thickness changes become too significant to be
neglected.
In this paper, a new testing technique has been adopted following EM 1110-2-1906 (USACE,
1970), EM 1110-2-5027 (USACE 1987) Appendix D, ASTM D-2435 Method B (1996),
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Louisiana Coastal Protection and Restoration Authority Geotechnical Standards (CPRA 2017),
and the method proposed by Azimi (2018). As a partial modification of the conventional
oedometer test, a 3D printing technology was applied to produce a prototype of the stainless-
steel dial cap using lightweight materials. Instead of using the stainless steel dial cap, the 3D
printed dial cap reduces the seating pressure dramatically. A series of consolidation tests were
performed for Louisiana marsh soils using both conventional and modified oedometers. Methods
and results of the consolidation tests are presented in this paper. The seating load is 0.01 TSF for
the conventional stainless steel dial cap and 0.002 TSF for the 3D printed one, and a new loading
schedule was developed for the modified oedometer test. The existing linear and the nonlinear
equations for the relationship between void ratio and effective stress, recommended by U.S.
Army Corps of Engineers (USACE), were used to identify the validity of the results presented in
this study.

THE DREDGED SOIL PROPERTIES AND SLURRY SAMPLE PREPARATIONS

The slurry soils were prepared using the drilled soil samples collected from the No- Name
Bayou Swamp production site in Louisiana, reflecting the dredged soils' natural condition.
Before slurry soil samples were prepared for consolidation tests, basic investigations were
carried out. The dredged soil samples are classified as Low Plastic Clay (CL) based on the
Unified Soil Classification System ASTM D2487 (1984). The measured specific gravity GS
according to ASTM D854 (ASTM 2016) was 2.6. The Atterberg limits were measured according
to the ASTM D4318 (2018) standard, and are shown in Figure 1(a). The soils were first oven-
dried for 24 hours, as shown in Figure 1(b) and then crushed with a grinder. The dry soil mass
was used as base material for the preparation of the slurry samples. After water was added to the
crushed dry soil to produce a homogenous slurry sample with an expected moisture content, the
wet soil sample was moved into a mixing cup for mixing for about two minutes until it became a
homogeneous slurry. The slurry samples came up with the moisture contents of 210.1%, 180.3%,
and 150.2%, respectively, as shown in Figure 1(c). To mimic the natural condition and lower the
initial moisture content before conducting the oedometer consolidation tests, the slurry samples
were poured into an acrylic settling column cylinder with a height of 10 inches. The soil particles
were first settled down by gravity, and the void ratio of the slurry soil decreased as the pore
water dispersed and gathered on the top of the settled soil in the settling column. It took almost
four weeks to complete the self-weight consolidation in each test. For the slurry samples with the
moisture content of (210.1%, 180.3% and 150.2%), after the settling column cylinder tests were
done, the average moisture content at the top of the settled soils was 150%, and decreased to
60% at the bottom of the settling column cylinder. Upon the completion of each settling column
test, as the first step, water at the top of the settling cylinder was removed using a hand pump,
and then the soil specimens were collected from different depths for performing oedometer
consolidation tests. To determine the soil specimen’s moisture content, some portions of the soils
at the same depth were collected and tested according to ASTM D2216 (2019).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 333

80
Plasticity Index (%) (a)
60
U-line
40
A-line
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

20

0
0 20 40 60 80 100
Liquid Limit (%)

Figure 1. (a) Atterberg limits of soil samples from the No-Name Bayou marsh creation site
at Louisiana, (b) Oven-dry soil mass, (c) Prepared slurry sample for laboratory
consolidation tests.

CONSOLIDATION TESTS USING CONVENTIONAL AND MODIFIED


OEDOMETERS

To reduce the seating pressure from the traditional stainless steel dial cap in any conventional
consolidation test, 3-D printing technology was applied to make up an identical dial cap with
exactly the same shape and sizes but a much lighter weight. Instead of using regular steel
material, a special alloy called Polylactic Acid (PLA) was employed. As presented in Figure 2,
the 3D-printed PLA dial cap weighs only 20.2 grams, while the conventional stainless-steel dial
cap weighs 387.5 grams. After applying a load of 2 TSF, the deflection of the 3D printed PLA
dial cap and the conventional stainless-steel dial cap were measured at 0.026 mm and 0.002 mm
respectively. As another way to check the measured deflections of the dial caps, finite element-
based software SolidWorks was applied. Numerical results with the finite element analyses came
up with 0.014 mm and 0.001 mm, respectively, using the Solid-Works software, as shown in
Figure 3. For the slurry consolidation tests, a 24-hour consolidation settlement could reach 50%
of the original sample depth (25.4 mm) at 1 TSF, so the error was less than about 0.2%.
According to literature review, the elastic modulus of 3500 MPa, yield strength of 70 MPa, and
the Poisson’s ratio of 0.36 were taken for the PLA materials (Lay et al., 2019, Torres et al., 2015)
in the finite element analyses. For stainless steel default mechanical properties in SolidWorks
software were used. The inner diameter of the oedometer dial ring is 63.5 mm, and the height is
25.4 mm.
In the modified 1-D oedometer test, the light-weight 3D printed dial cap has achieved a
seating pressure of about 0.002 TSF, including the pressure that comes up from the weight of
the porous stone, as opposed to the traditional 0.01 TSF. The testing results indicated that the
maximum initial water content of the slurry soil sample was unable to go beyond 70% if the
conventional stainless steel dial cap was used. In some cases, soil samples started leaking to
the specimen chamber if the moisture content was 60% or higher. It might conclude that the
slurry soil specimens with a moisture content equal to or greater than 70%, the consolidation
tests with the regular stainless-steel dial cap would fail. However, if the 3D-printed dial cap
was used, the maximum initial water content of the slurry specimens could go up as high as
100%.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 334
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. A stainless-steel dial cap (left) and a 3D printed polylactic acid (PLA) dial cap
(right): (a) Top view, (b) Bottom view, (c) The stainless-steel dial cap with a weigh of 387.5
grams, (d) The 3D printed dial cap with a weigh of 20.2 grams, (e) A consolidation test in-
progress using the stainless-steel dial cap. (f) A consolidation test in progress using the 3D-
printed dial cap.

*UZ = Deflection in the Z direction.

Figure 3. Deflections of dial caps subjected to the vertical load of 2 TSF using the
SolidWorks software: (a) PLA, (b) Stainless steel.

During the successful consolidation tests using the 3D printed dial cap, the following loading
schedule was adopted: 0.002, 0.005, 0.01, 0.025, 0.05, 0.10, 0.25, 0.50, and 1.00 TSF. The 1-D
consolidation tests were performed according to ASTM D-2435 (ASTM 1996) and EM 1110-2-
5027 (USACE 1987) Appendix D. Consolidation readings were taken at times 0.1, 0.2, 0.5, 1.0,
2.0, 4.0, 8.0, 15.0 and 30.0 minutes, 1, 2, 4, 8 and 24 hours. During the 24-hour time period,
100% of the primary consolidation was complete for most of the applied load mentioned above.
The consolidation test readings were analyzed following Method B of ASTM D-2435 (ASTM
1996) and results were presented in the analyses of the consolidation test results section in this
paper.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 335
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Laboratory test curves for the relationship between void ratio and the effective
stress for soil specimens with different moisture contents (a) using the conventional
oedometer with stainless-steel dial cap, (b) using the modified oedometer with 3D printed
dial cap.

ANALYSES OF THE CONSOLIDATION TEST RESULTS

Determination of the consolidation properties was done following Method B of standard


ASTM 2435 (ASTM 1996). A total of six samples were tested for the conventional consolidation
test, and the modified consolidation test, respectively. The soil specimen containing 68.1%
moisture was successfully tested using a conventional stainless-steel dial cap. Another two soil
specimens were tested, which had moisture contents of 65.2% and 61.4%, respectively. The 3D
printed dial caps, which came up with a seating load of 0.002 TSF, including the porous stone's
weight, could successfully test the soil specimen with a moisture content of 101.3 %. The other
two soil specimens were successfully tested, having moisture contents of 92.9% and 82.4%,
respectively.
The void ratio versus effective stress curves of conventional oedometer tests using stainless
steel dial caps and modified oedometer tests using 3D printed dial caps are presented in Figure 4.
All the curves showed similar trends, and the void ratios of all the six soil specimens turned out
almost the same results, which were below 1 after the application of the load of 1 TSF. After
applying the 1 TSF load, the vertical strains of all the tested specimens reached nearly 50%,
indicating that a large consolidation deformation happened. And thus, no further loading was
imposed to continue the consolidation tests. According to Apu et al. (2021), Sarker et al. (2021)
and Apu & Wang (2022) the compression index (𝒄𝒄 ) of Louisiana Marsh soil would be between
0.16 - 2.86. From the lab experiments of these six soil specimens, the compression indices fell in
the range between 0.33 - 0.48, which are presented in Figure 6. The values were seemingly
reasonable.
The measured coefficients of consolidation 𝑐𝑣 from the conventional oedometer tests ranged
from 9.0*10-9 m2/s to 9.7*10-8 m2/s (Figure 5a-5c). and the values from the modified oedometer
tests ranged from 1.0*10-8 m2/s to 9.6*10-8 m2/s (Figure 5d-5f). It is observed that value of 𝑐𝑣
decreases with the increasing effective stress and after increasing the load to 0.01 TSF, 𝑐𝑣
decreases significantly. Holtz (1972) reported 𝑐𝑣 values ranging from 4*10-9 m2/s to 7*10-9 m2/s
for clay from Mexico City which is almost similar to the dredged Louisiana soil. In contrast,
measured 𝑐𝑣 values for Ottawa sand mixed with 15% silt ranged from 4.4*10-4 m2/s to 1.2*10-3

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 336

m2/s when consolidated in triaxial tests and from 6.2*10-4 m2/s to 2.7*10-3 m2/s when
consolidated in oedometer tests (Carraro et al., 2003). The 𝑐𝑣 values for foundry sand mixed with
15% non-plastic silt ranged from 3.5*10-4 m2/s to 1.7*10-3 m2/s (Thevanayagam et al. 2001).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5. Effective stress vs. Coefficient of consolidation (𝒄𝒗 ) and Hydraulic conductivity
(k) of the soil specimens with different moisture contents: (a - c) Using the conventional
oedometer with a stainless steel dial cap, (d -f) Using the modified oedometer with a 3D
printed dial cap.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 337

The measured hydraulic conductivity k values from the convention oedometer tests were in
the range from 5.2*10-11 m/s to 6.5*10-8 m/s (Figure 5a-5c) and values from the modified
oedometer tests ranged from 1.1*10-9 m/s to 3 .5*10-8 m/s (Figure 5d-5f). The hydraulic
conductivity of the Louisiana marsh soil samples decreased significantly in both conventional
and modified oedometer tests after applying a load of 0.01 TSF. Li et al. (2013) measured the k
values of highly lean clay ranging from 3.3*10-7 m/s to 1.2*10-6 m/s, and Moozhikkal et al.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(2019) measured the k values of Bombay marine clay ranging from 1.0*10-10 m/s to 1.9*10-8 m/s,
which were similar to the measured values of k for slurries from the Louisiana dredge soils
presented in this study.

Figure 6. Laboratory test curves for the relationship between Effective stress and Void
ratio with best apparent fit curve fitted using equations (1) & (2): (a) through (c) using
conventional stainless-steel dial cap, (d) through (f) using 3D printed dial cap.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 338

VALIDATION USING USACE RECOMMENDED EQUATION

USACE recommended Equation (1) for the relationship between void ratio and effective
stress for large strain consolidations, initially proposed by Gibson et al. (1981) following the
nonlinear consolidation theory. Gibson et al. (1981) concluded that the form of Equation (1) is
compatible with the general form of the relationship between void ratio and effective stress
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

found in conventional soils, where 𝑒𝑜 is the voids ratio at zero effective stress and 𝑒∞ is the voids
ratio at the end of consolidation and λ is the variable coefficient. In addition, Gibson et al. (1981)
presented Equation (1) as a replacement for Equation (2), commonly used to describe oedometer
test results for normally consolidated soil, where 𝑒𝑜 and 𝜎𝑜′ are the reference void ratio and
effective stress, respectively, and 𝑐𝑐 is the compression index. To determine the validity of the
test and analysis, the void ratios from the lab tests were fitted to Equations (1) and (2) against the
void ratio vs. effective stress curves obtained by the experiment and presented in Figure 6.

𝑒 = (𝑒𝑜 − 𝑒∞ )𝐸𝑋𝑃(−λ𝜎 ′ ) + 𝑒∞ (1)

𝜎′
𝑒 = 𝑒𝑜 − 𝑐𝑐 𝑙𝑜𝑔10 ( ′ ) (2)
𝜎𝑜

To determine the validity of the test and analysis method described in this study, the void
ratio vs. effective stress curves were compared with the plots created from the USACE
recommended equation, as presented in Figure 6 (a – f). The void ratio vs. effective stress curves
from all the tests were fitted with non-linear equation (1) and linear equation (2), respectively,
and the corresponding values of e∞, e0, λ, and 𝑐𝑐 for both equations were noted in each figure.
From the analyses and the established equations from curve-fitting, it is observed that the
experimental data follows almost the same trend. The non-linear curve-fitting equations
following equation (1) are more accurate than those following linear equation (2). From these
observations, it may say that the marsh soil consolidation process might be more properly
described using the non-linear consolidation theory, which would give a more precise settlement
prediction.

CONCLUSION

A conventional oedometer test cannot handle very soft soils with a high-water content. As an
alternative, 3D printing technology could be a useful and economical tool to modify the
conventional oedometer to handle very soft mud with a high moisture content. The conclusions
of this study are listed below:
1. Using the lightweight 3D printed dial cap instead of the conventional stainless steel one,
the seating load that combines the weights of dial cap and porous stone decreases more
than ten times.
2. The conventional oedometer may not handle the soil with more than 70% moisture
content, whereas the modified oedometer using 3D printing technology may handle soil
specimens with a moisture content up to 100%.
3. Comparing the coefficients of consolidation and the hydraulic conductivities of the
Louisiana marsh soil in this study with the results from previous studies, it may consider
that results from the modified oedometer tests fell in a reasonable range.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 339

4. The existing non-linear consolidation theory could be utilized to describe the


consolidation of marsh soils with high moisture contents. The non-linear consolidation
theory is more accurate than the linear theory.

ACKNOWLEDGEMENTS
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

The research presented in this paper was funded by Louisiana Sea Grant. The grant comes up
originally from the Coastal Protection and Restoration Authority of Louisiana (CPRA). We are
thankful to Dr. David Hall and Dr. Kelly Crittenden for their support and assistance in drawing
the dial cap in Solid-Works and printing out the dial caps using the 3D printers in the College of
Engineering, Louisiana Tech University. Support and assistance from Mr. Jacques Boudreaux
and Mr. Russ Joffrion are gratefully acknowledged.

REFERENCES

Apu, O. S., and Wang, J. X. (n.d.). Assessment of Compression Index (cc) of Louisiana Marsh
Soils by Considering the Sedimentation State. Geo-Congress 2022, 131–140.
Apu, O. S., Wang, J. X., ad Sarker, D. (2021). Evolution of Large-Strain One-Dimensional
Consolidation Test for Louisiana Marsh Soil. In IFCEE 2021 (pp. 244–255).
ASTM. (2016). Standard Test Methods for Specific Gravity of Soil Solids by Water
PycnometerNo Title. ASTM.
ASTM. (2018). Standard Test Methods for Liquid Limit, Plastic Limit, and Plasticity Index of
Soils. ASTM.
ASTM. (2019). Standard Test Methods for Laboratory Determination of Water (Moisture)
Content of Soil and Rock by Mass. ASTM.
ASTM. (1996). Standard test method for one-dimensional consolidation properties of soils.
Azimi, A. (2018). Laboratory Self-Weight Consolidation Testing of Dredged Material Analyzed
Using a One-Dimensional Finite Strain Consolidation Method. Master of Science in Civil
Engineering Theses., Kennesaw State University,
GA,htps://digitalcommons.kennesaw.edu/msce_etd/2.
Carraro, J. A. H., Bandini, P., and Salgado, R. (2003). Liquefaction resistance of clean and
nonplastic silty sands based on cone penetration resistance. Journal of Geotechnical and
Geoenvironmental Engineering, 129(11), 965–976.
CPRA (Coastal Protection and Restaration Authority). (2017). Geotechnical Standards Marsh
Creation and Coastal Restoration Projects. Coastal Protection and Restaration Authority of
Louisiana.1–45.
Gibson, R. E., Schiffman, R. L., and Cargill, K. W. (1981). The theory of one-dimensional
consolidation of saturated clays. II. Finite nonlinear consolidation of thick homogeneous
layers. Canadian Geotechnical Journal, 18(2), 280–293.
Holtz, R. D. (1972). Long-term loading tests at Ska-Edeby, Sweden. Proceedings ASCE
Specialty Conference on Earth and Earth Supported Structures, 435–464.
Howard, A. K. (1984). The revised ASTM standard on the unified classification system.
Geotechnical Testing Journal, 7(4), 216–222.
Imai, G. (1978). Fundamental studies on one-dimensional consolidation characteristics of fluid
mud. Doctoral Dissertation, University of Tokyo (in Japanese).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 340

Lay, M., Thajudin, N. L. N., Hamid, Z. A. A., Rusli, A., Abdullah, M. K., and Shuib, R. K.
(2019). Comparison of physical and mechanical properties of PLA, ABS and nylon 6
fabricated using fused deposition modeling and injection molding. Composites Part B:
Engineering, 176, 107341.
Li, L., Alvarez, I. C., and Aubertin, J. D. (2013). Self-weight consolidation of slurried
deposition: tests and interpretation. International Journal of Geotechnical Engineering, 7(2),
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

205–213.
Mikasa, M. (1963). The consolidation of soft clay-a new consolidation theory and its
application.
Moozhikkal, R., Sridhar, G., and Robinson, R. G. (2019). Constant rate of strain consolidation
test using conventional fixed ring consolidation cell. Indian Geotechnical Journal, 49(2),
141–150.
Sarker, D., Shahrear Apu, O., Kumar, N., Wang, J. X., and Lynam, J. G. (2021). Application of
sustainable lignin stabilized expansive soils in highway subgrade. In IFCEE 2021 (pp. 336–
348).
Thevanayagam, S., Martin, G. R., Shenthan, T., and Liang, J. (2001). Post-liquefaction pore
pressure dissipation and densification in silty soils.
Torres, J., Cotelo, J., Karl, J., and Gordon, A. P. (2015). Mechanical property optimization of
FDM PLA in shear with multiple objectives. Jom, 67(5), 1183–1193.
Umehara, Y. (1975). Determination of consolidation constants for very soft clay. Report of The
Port and Harbour Research Institute, 14(4), 45–65.
Umehara, Y., and Zen, K. (1980). Constant rate of strain consolidation for very soft clayey soils.
Soils and Foundations, 20(2), 79–95.
USACE. (1970). Laboratory soils testing. EM 1110-2–1906. Washington, DC: US Army Corps
of Engineers.
USACE. (1987). Confined disposal of dredged material.EM 1110-2-5027. Washington, DC: US
Army Corps of Engineers.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 341

Effect of Thermal Volume Change on the Permeability of Kaolin Clay during a Heating-
Cooling Cycle

Nilufar Chowdhury, S.M.ASCE1; and Omid Ghasemi-Fare, Ph.D., A.M.ASCE2


1
Ph.D. Student, Dept. of Civil and Environmental Engineering, Univ. of Louisville, Louisville.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Email: n0chow03@louisville.edu
2
Associate Professor, Dept. of Civil and Environmental Engineering, Univ. of Louisville,
Louisville. Email: omid.ghasemifare@louisville.edu

ABSTRACT

Thermal loading changes the hydraulic conductivity of the soil medium. This research aims
to study the effect of a heating-cooling cycle on the permeability of Kaolin clay. A temperature-
controlled triaxial permeameter cell is used in this study to explore the variations of soil
permeability at 20°C, 50°C, and 80°C. The hydraulic conductivity at each temperature is
measured after 48 h at each temperature during the heating-cooling cycle to make sure the
volume change due to thermal loading is stabilized. Evaluation of void ratio change during
heating and cooling has been observed to predict the thermal behavior of Kaolin clay. It is
observed that the intrinsic permeability during the cooling cycle is 9%–12% lower than the
permeability observed during the heating cycle. The obtained results suggest that Kaolin clay’s
microstructure can change subject to a heating-cooling process, which creates irrecoverable
deformation in the clay and regulates the macro behavior such as permeability of Kaolin clay.

INTRODUCTION

Permeability is one of the most significant engineering properties of clay that regulates the
seepage and water movement in the ground. Hydraulic properties such as permeability
emphasize many geotechnical challenges, such as slope stability analysis and landfill designs
(Damiano et al., 2017; Dou et al., 2014; Ng and Leung, 2012; Sadeghi and AliPanahi, 2020). The
engineering properties of fine-grained soils are subjected to changes in external pressure, wetting
and drying cycles, the chemistry of the pore medium, and temperature (Sridharan, 2002).
Therefore, previous researchers have investigated the temperature effect on the volumetric strain
of Normally Consolidated (NC) clay and over-consolidated (OC) clay under the heating-cooling
cycle (Baldi et al. 1988; Samarakoon & McCartney, 2020). It is observed that due to the drained
heating and cooling, respectively, plastic volumetric contraction, and elastic contraction may
occur in normally consolidated soil (Samarakoon & McCartney, 2020). Moreover, temperature
affects the soil structure, inter-particle bond strength, void ratio, water holding capacity of
particles, and the corresponding hydraulic properties of soil (Gao and Shao, 2015).
A significant component of the thermomechanical behavior of saturated clay at elevated
temperatures is the volume change of components of a clay mineral-water system due to the
thermal effect (Baldi et al. 1988). According to Cekerevac and Laloui, (2004), soils with highly
over consolidation ratio (OCRs) show reversible thermal expansion followed by irreversible
thermal contraction. On the other hand, normally consolidated soils show irreversible volumetric
contraction while heated at slow heating rates. This occurs because of the decrease in inter-
particle shearing strength during the thermal load causing a collapse of the soil skeleton,

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 342

eventually reducing the void ratio of the specimen (Campanella and Mitchell, 1968). The
aforementioned thermal strains in saturated clays result from the water behavior, thermal
expansion of clay minerals, reordering of the clay skeleton structure, as well as drainage
conditions (Campanella and Mitchell 1968).
Thermal loading alters soil and fluid properties (Cherati and Ghasemi-Fare, 2019; Monfared
et al., 2014; Tamizdoust and Ghasemi Fare, 2020a). Therefore, to properly understand the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

thermo-hydro-mechanical response of the geotechnical infrastructures, the variations of the soil


properties (i.e., permeability) with the temperature change must be studied (Tamizdoust and
Ghasemi-Fare, 2020b; Chen et al., 2017; François et al., 2009; Ghasemi-Fare and Basu, 2019).
Previous research reported different behaviors for permeability variations of different clays at
elevated temperatures. Towhata et al. (1993) and Seiphoori (2015) observed an increase in
permeability of clay while some researchers have observed a decrease in intrinsic permeability at
elevated temperatures due to the thermal consolidation induced soil densification (Romero et al.
2001; Villar and Lloret, 2004). Alternatively, Delage et al. (2009) conducted temperature-
controlled constant head permeability tests on Boom clay and reported no changes in intrinsic
permeability in the 20 °C to 90 °C temperature range. Chen et al. (2017) investigated the effect
of thermal loading and the corresponding microstructural change on the hydraulic conductivity
of Boom clay. They have reported that the microstructure weakening due to the thermal volume
change resulted in intrinsic permeability reduction in Boom clay. Similarly, Joshaghani and
Ghasemi-Fare (2021) have reported although hydraulic conductivity increases with temperature
for Kaolin clay, the intrinsic permeability of Kaolin clay slightly reduces when the temperature
rises from 20 °C to 80 °C.
However, in these previous studies, the focus was given to the changes in hydraulic
conductivity as well as the intrinsic permeability of soil during the heating/ cooling cycle. It still
remains unclear how thermal volume changes and induced soil fabric alteration during a heating-
cooling cycle affect clay's hydraulic conductivity at a particular temperature. This study aims to
analyze the hydraulic conductivity, intrinsic permeability, as well as thermal volume change of a
normally consolidated (NC) Kaolin clay during a heating-cooling cycle. The results help to
quantify and compare the thermal volume change (void ratio) during heating and cooling and its
effect on hydraulic conductivity and intrinsic permeability during a heating and cooling load

EXPERIMENTAL PROGRAM

Sample Preparation

Commercial natural Kaolin clay from Columbus Co. has been selected for this study. To
prepare the specimen, the Kaolinite powder was hand mixed with de-ionized water having an
initial water content of 30 %. The soil was mixed with water carefully so that no soil lumps exist.
Then the soil mixture was compacted into a cylindrical specimen with a height of 25.4 mm, a
diameter of 101 mm, and a dry density of 1.4 Mg/m3. Compaction was done in three layers. In
order to maintain the same void ratio for the specimens, the height of the specimen was carefully
controlled. After compaction, the height of each specimen was determined, and the initial void
ratios of all specimens were kept almost constant. Then, the compacted sample was extruded
from the cylindrical ring. After that, the specimen was placed between the filter papers and
porous stones. Then the membrane was placed around it. The hydraulic conductivity was
performed at higher confinement (stress) compared to the initial stress state; thus, all clayey

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 343

samples were considered normally consolidated clays. A modified triaxial cell capable of
controlling the heating-cooling cycle (Joshaghani and Ghasemi-Fare, 2021) was used in this
study. The triaxial system was calibrated for thermal deflections (Joshaghani and Ghasemi-Fare,
2020, 2021). The cell was filled with de-aired water, as well as the sample was saturated by
allowing the de-aired water to pass through the specimens by applying a pressure gradient
between the bottom and the top of the sample. The saturation of the sample was controlled
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

before beginning each test by calculating Skempton's coefficient B value. Cell pressure was
adjusted and controlled by one of the digitally controlled flow pumps of the triaxial setup.
The geotechnical and thermal properties of the Kaolin clay used in this research are presented
in Table 1. The thermal variations of viscosity and density of “pure water” are adopted from
Delage et al. (2009).

Table 1. Properties of Kaolin clay.

Liquid Plasticity Specific Specific VCL RCL Thermal Specific


Limit Index Gravity Surface area Slope Slope conductivity heat
(λ) (κ) W/(m·K) capacity of
(%) (m²/g) Kaolinite;
Cp
(J/(kg◦C))
40 10 2.63 12.7373 0.228 0.08 1.798 945 (at 20
◦C)

Test procedure

To perform hydraulic conductivity tests at different temperatures, the adopted test setup was
similar to Joshaghani and Ghasemi-Fare, (2021). The detailed description of the modified
temperature-controlled triaxial cell can be found in Joshaghani and Ghasemi-Fare, (2021).
After achieving the approximate full saturation of the sample, the pressure gradient has been
applied at the bottom and top of the sample. The applied hydraulic gradient was 138". To
perform a test, initially, the temperature of the cell was fixed to the desired temperature using a
temperature-controlled water bath. The bottom drain valve of the cell has been kept open to
allow the generated pore water pressure to dissipate. The volume of dissipated water has been
measured by the digital flow pump. Then, the hydraulic conductivity test has been started by
employing a pressure gradient between the top and bottom of the sample. The volume of the
passed water during the test was measured using a flow pump. After measuring the hydraulic
conductivity (H.C) at room temperature (20 ºC), the sample temperature increased to 50 ºC. The
hydraulic conductivity and the void ratio changes during a single heating-cooling test are
measured after 48 hours of thermal loading. Please note, that 48 hours were selected such that
after which no observable volume changes occurred, indicating stabilization of thermal volume
change due to thermal loading. At every temperature, H.C is measured three times to ensure the
repeatability of the test. To calculate the thermal volume change of the specimen, the equation
((∆Vdr)∆T= αwVw∆T + αsVs∆T - (∆Vm)∆T), proposed by Campanella and Mitchell, (1968) has
been adopted where the thermal volume change of the specimen (∆Vm) is predicted by
measuring the drained or absorbed water (∆Vdr) while taking into account of thermal expansion
of water (αwVw∆T) and soil particles (αsVs∆T), respectively. Intrinsic permeability is calculated

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 344

ˠ𝑤
considering the density and viscosity change of water with temperature (ĸ = 𝐾 ,where ŋ is
ŋ
dynamic viscosity, ˠ𝑤 is the unit weight of water and 𝐾 is intrinsic permeability, ĸ is the
hydraulic conductivity). Change in the void ratio is also obtained by calculating Vw and Vs after
each thermal load.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

RESULTS

Figure 1 illustrates the change of the void ratio of Kaolin clay with a heating-cooling cycle to
identify the effect of the thermal volume change of NC Kaolin clay on its hydraulic conductivity
as well as intrinsic permeability. It is observed that the void ratio during the heating cycle
decreased by 21% while the temperature increased from 20°C to 80°C.
This reflects that increased temperature causes particle rearrangement, resulting in thermal
volume contraction of normally consolidated Kaolin clay.

0.8

0.7
Void Ratio (e)

0.6

0.5

0.4

0.3 Heating Cycle


Cooling Cycle
0.2
0 20 40 60 80 100
Temperature (°C)

Figure 1. Changes in the void ratio of Kaolin clay with the heating-cooling cycle under 552
kPa confinement stresses

It is interesting to report that the void ratio increased during the cooling cycle by 7% while
the temperature decreased from 80°C to 20°C. The changes in void ratio during the cooling load
also confirm the soil particle rearrangement happens due to the cooling load. The 7% thermal
volume expansion confirms that a part of the volume reduction that happens during the heating
load was elastic and reversible.
Table 2 represents the differences in the void ratio between the heating and cooling cycle. It
is observed that there is an irrecoverable change in the void ratio of Kaolin clay after a complete
heating-cooling cycle. Under 552 kPa confining pressure, during heating, the void ratio of the
sample is 0.69, 0.64, and 0.54 at 20°C, 50°C, and 80°C, respectively.
But during cooling, the void ratio of the sample is 0.58, and 0.546 at 20°C and 50°C,
respectively, showing almost 16% and 15% of irrecoverable volume contraction at 20°C and
50°C, respectively. This result reflects that, during a heating and cooling cycle, there is always a

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 345

difference in void ratio at a particular temperature, reflecting the irrecoverable volumetric change
in the sample due to the thermal load.

Table 2. Differences in void ratio of Kaolin clay between the heating and cooling cycle
under 552 kPa confinement stresses
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Initial Void Ratio


Confining Void Ratio (e) Void Ratio (e) Reduction (after
Temperature (°C)
Pressure (after heating) (after Cooling) heating and cooling
(kPa) cycle) (%)
552 20 0.69 0.58 16
50 0.64 0.546 15
80 0.54 0.54 -

Figure 2 represents the changes in the hydraulic conductivity of kaolin clay during the
thermal heating load. It is evident that hydraulic conductivity increases significantly when the
temperature changes from 20 ºC to 50 ºC and then to 80 ºC. Under 552 kPa confining pressure,
with the temperature increasing from 20ºC to 80ºC, the hydraulic conductivity increases by
113%. The increase in H.C. at higher temperatures occurs due to the reduction of the dynamic
viscosity of water at elevated temperatures, which can facilitate water movements in the flow
channels.

1E-09
Hydraulic Conductivity (m/s)

8E-10

6E-10

4E-10

2E-10

0
20 50 80
Temperature (°C)

Figure 2. Variations in hydraulic conductivity of Kaolin clay with the heating cycle under
552 kPa confinement stress

Figure 3 demonstrates the changes in the hydraulic conductivity of Kaolin clay during the
cooling cycle. It is observed that hydraulic conductivity decreases significantly when the
temperature reduces from 80 ºC to 50 ºC and then to 20 ºC. Under 552 kPa confining pressure,
when the temperature decreases from 80 ºC to 20 ºC, the hydraulic conductivity decreases by
58%.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 346

1E-09

Hydraulic Conductivity (m/s)


8E-10

6E-10
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

4E-10

2E-10

0
80 50 20
Temperature (°C)

Figure 3. Variations in hydraulic conductivity of Kaolin clay with the cooling cycle under
552 kPa confinement stress

Table 3 represents the differences in H.C value observed during the heating and cooling
loads. It is observed that there is a permanent change in the H.C of Kaolin clay after a complete
heating-cooling cycle. For instance, under 552 kPa confining pressure, during heating, the H.C of
the sample is 3.97E-10 m/s, 6.40E-10 m/s, and 8.46E-10 m/s at 20 ºC, 50 ºC, and 80 ºC,
respectively. On the other hand, during cooling from 80 ºC to 20 ºC, the H.C of the sample is
3.55E-10 m/s, 6.06E-10 m/s at 20 ºC and 50 ºC, respectively, denoting almost 11% and 5% of
lower H.C values, respectively, at 20 ºC and 50 ºC, after the cooling load is applied compared the
values observed during the thermal loading.
This result reflects that, during a heating and cooling cycle, at a particular temperature, there
is always a difference in H.C values of Kaolin clay, indicating irrecoverable thermal volume
change in soil fabric due to the thermal load.

Table 3. Differences in hydraulic conductivity of Kaolin clay between the heating and
cooling cycle under 552 kPa confinement stress

Initial Hydraulic Hydraulic


Hydraulic
Confining Conductivity, Conductivity
Temperature (°C) Conductivity, m/s
Pressure m/s (during Differences
(during Cooling)
(kPa) heating) (%)
552 20 3.97E-10 3.55E-10 11
50 6.40E-10 6.06E-10 5
80 8.46E-10 8.46E-10 -

Figure 4 represents the changes in the intrinsic permeability of Kaolin clay with a heating
load. As can be seen in Figure 5, the intrinsic permeability decreases when the temperature
increases from 20 ºC to 50 ºC and then to 80 ºC. Under 552 kPa confining pressure, with the
temperature increase from 20ºC to 80ºC, the intrinsic permeability decreases by 22%. The

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 347

decrease in intrinsic permeability occurs due to the thermal volume contraction in higher
temperatures.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Changes in Intrinsic Permeability of Kaolin clay with the heating cycle under 552
kPa confinement stress

Figure 5 demonstrates the changes in the intrinsic permeability of kaolin clay with a cooling
cycle. The results determine that intrinsic permeability increases when the temperature reduces
from 80 ºC to 50 ºC and then to 20 ºC. Under 552 kPa confining pressure, when the temperature
decreases from 80 ºC to 20 ºC, the intrinsic permeability increases by 15%.

5E-17
Intrinsic Permeability (m2)

4E-17

3E-17

2E-17

1E-17

0
80 50 20
Temperature (°C)

Figure 5. Changes in Intrinsic Permeability of Kaolin clay with the cooling cycle under 552
kPa confinement stress

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 348

Table 4 represents the differences in intrinsic permeability value between the heating and
cooling loads. It is observed that the intrinsic permeability of Kaolin clay measured after a
complete heating-cooling cycle is different from its initial value. For instance, at 552 kPa
confining pressure, during heating, the intrinsic permeability of the sample is 4.1E-17 m2, 3.71E-
17 m2, and 3.16E-17 m2 at 20 ºC, 50 ºC, and 80 ºC, respectively. But during cooling load when
soil temperatures are reduced from 80 ºC to 20 ºC, the intrinsic permeability of the sample is
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

3.62E-17 m2 and 3.38E-17 m2 at 20ºC and 50ºC, respectively, showing almost 12% and 9% of
lower intrinsic permeability values during cooling load compared to the values measured at 20ºC
and 50ºC, respectively, during the heating load.

Table 4. Differences in intrinsic permeability of Kaolin clay between the heating and
cooling cycle under 552 kPa confinement stress

Initial Intrinsic
Intrinsic Intrinsic
Confining Permeability,
Temperature (°C) Permeability, m2 Permeability
Pressure m2 (during
(during Cooling) Differences (%)
(kPa) heating)
552 20 4.1E-17 3.62E-17 12
50 3.71E-17 3.38E-17 9
80 3.16E-17 3.16E-17 -

This result reflects that, during a heating and cooling cycle, at a particular temperature, there
is always a difference in intrinsic permeability values of Kaolin clay, indicating a change in soil
fabric due to the thermal load.

CONCLUSIONS

The hydraulic conductivity of Kaolin clay under 552 kPa confining stresses is measured.
Evaluation of void ratio (thermal volume change) during a heating and cooling cycle has been
observed to predict the thermal volume change of Kaolin clay. The following conclusions can be
illustrated based on the results of this study:
1. After a complete heating-cooling cycle, there is a change in the void ratio compared to
the sample's initial void ratio, indicating an irrecoverable volumetric contraction of
Kaolin clay due to the thermal load. Hydraulic conductivity measured during a heating
model is reduced by 5 to 11% compared to the permeability observed prior to a thermal
load. This reduction in hydraulic conductivity is attributed to the irreversible thermal
volume change of NC Kaolin clay.
2. Intrinsic permeability during the cooling cycle is 9 to 12% lower compared to the
permeability observed during the heating cycle at 20°C temperature.
The obtained results suggest that Kaolin clay’s fabric can change when it is subjected to a
complete heating-cooling load which creates thermal deformation in the clay and regulates the
macro behavior such as the permeability of Kaolin clay.

ACKNOWLEDGEMENT

The authors would also like to gratefully acknowledge the financial support by the National
Science Foundation under Grant No. CMMI-1804822.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 349

REFERENCES

Baldi, G., Hueckel, T., and Pellegrini, R. 1988. Thermal volume changes of the mineral–water
system in low-porosity clay soils. Canadian geotechnical journal, 25(4), 807-825.
Campanella, R. G., and Mitchell, J. K. 1968. Influence of temperature variations on soil
behavior. Journal of Soil Mechanics & Foundations Div.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Cekerevac, C., and Laloui, L. 2004. Experimental study of thermal effects on the mechanical
behaviour of a clay. International journal for numerical analytical methods in geomechanics,
28, 209-228.
Chen, W., Ma, Y., Yu, H., Li, F., Li, X., and Sillen, X. 2017. Effects of temperature and
thermally induced microstructure change on hydraulic conductivity of Boom Clay. J. Rock
Mech. Geotech. Eng. 9, 383–395.
Cherati, D. Y., and Ghasemi-Fare, O. J. G. 2019. Analyzing transient heat and moisture
transport surrounding a heat source in unsaturated porous media using the Green’s function,
81, pp. 224–234.
Damiano, E., Greco, R., Guida, A., Olivares, L., and Picarelli, L. 2017. Investigation on
rainwater infiltration into layered shallow covers in pyroclastic soils and its effect on slope
stability. Engineering Geology, 220.
Dou, H.-Q., Han, T.-C., Gong, X.-N., and Zhang, J. 2014. Probabilistic slope stability analysis
considering the variability of hydraulic conductivity under rainfall infiltration–redistribution
conditions. Engineering Geology, 183, 1-13.
Delage, P., Sultan, N., and Cui, Y. J. 2000. On the thermal consolidation of Boom clay. Can.
Geotech. J. 37, 343–354.
Delage, P., Sultan, N., Cui, Y.-J., and Ling, L. X. 2009. Permeability changes in Boom clay with
temperature. International Conference and Workshop “Impact of ThermoHydro-Mechanical-
Chemical (THMC) processes on the safety of underground radioactive waste repositories”,
European Union 331–335.
François, B., Laloui, L., and Laurent, C. 2009. Thermo-hydro-mechanical simulation of ATLAS
in situ large scale test in Boom Clay. Comput. Geotech. 36, 626–640.
Gao, H., and Shao, M. 2015. Effects of temperature changes on soil hydraulic properties. Soil
andTillage Research, 153, 145-154.
Ghasemi-Fare, O., and Basu, P. 2019. Coupling heat and buoyant fluid flow for thermal
performance assessment of geothermal piles. Comput. Geotech. 116, 103211.
Joshaghani, M., Ghavami, M., and Ghasemi-Fare, O. 2018. “Experimental investigation on the
effects of temperature on physical properties of sandy soils”, IFCEE, ASCE, March 5-10.
Joshaghani, M., and Ghasemi-Fare, O. 2019. “A study on thermal consolidation of fine grained
soils using modified triaxial cell”, Geo-congress, ASCE, March 24-27.
Joshaghani, M., and Ghasemi-Fare, O. 2021. Exploring the effects of temperature on intrinsic
permeability and void ratio alteration through temperature-controlled experiments.
Engineering Geology, 293, 106299.
Joshaghani, M., and Ghasemi-Fare, O. 2021. “Experimental study to analyze the effect of
confinement and cell pressure on thermal pressurization under fully undrained condition”,
International Foundations Congress and Equipment Expo, ASCE-GI, Dallas, TX, May 10-
14.
Monfared, M., Sulem, J., Delage, P., and Mohajerani, M., 2014. Temperature and damage impact
on the permeability of Opalinus clay. Rock Mech. Rock. Eng. 47, 101–110.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 350

Ng, C. W. W., and Leung, A. K. 2012. In-situ and laboratory investigations of stress-
dependentpermeability function and SDSWCC from an unsaturated soil slope. Geotech Eng,
43, 26-39.
Romero, E., Gens, A., and Lloret, A. 2001. Temperature effects on the hydraulic behaviour of an
unsaturated clay. In: Unsaturated Soil Concepts and Their Application in Geotechnical
Practice. Springer, pp. 311–332.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Sridharan, A. 2002. Engineering behaviour of clays: influence of mineralogy. Chemomechanical


coupling in clays, 3-28.
Samarakoon, R., and McCartney, J. S. 2020. Role of initial effective stress on the thermal
volume change of normally consolidated clay. In E3S Web of Conferences (Vol. 205, p.
09001). EDP Sciences.
Sadeghi, H., and AliPanahi, P. 2020. Saturated hydraulic conductivity of problematic soils
measured by a newly developed low-compliance triaxial permeameter. Engineering Geology,
105827.
Seiphoori, A. 2015. Thermo-Hydro-Mechanical Characterisation and Modelling of Wyoming
Granular Bentonite. Nagra.
Tamizdoust, M. M., and Ghasemi-Fare, O. 2020b. Coupled thermo-hydro-mechanical modeling
of saturated Boom clay. In: Geo-Congress 2020: Geo-Systems.
Tamizdoust, M. M., and Ghasemi-Fare, O. 2020a. Comparison of thermo-poroelastic and
thermo-poroelastoplastic constitutive models to analyze THM process in clays. In: E3S Web
of Conferences. EDP Sciences, p. 04008.
Towhata, I., Kuntiwattanaku, P., Seko, I., and Ohishi, K. 1993. Volume change of clays induced
by heating as observed in consolidation tests. Soils Found. 33, 170–183.
Villar, M. V., and Lloret, A. J. A. C. S. 2004. Influence of temperature on the hydro-mechanical
behaviour of a compacted bentonite, 26, pp. 337–350.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 351

Influence of Time-Dependent Soil Thermal Conductivity on Performance Assessment of


Energy Foundations

Arjun Sivaprasad, S.M.ASCE1; and Prasenjit Basu, M.ASCE2


1
Doctoral Research Fellow, Dept. of Civil Engineering, IIT Bombay, Mumbai.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Email: arjun_s@iitb.ac.in
2
Associate Professor, Dept. of Civil Engineering, IIT Bombay, Mumbai.
Email: pbasu@civil.iitb.ac.in

ABSTRACT

Soil thermal conductivity k plays a significant role in performance assessment of energy


geostructures that have direct or indirect thermal interaction with soil. Research presented herein
uses a custom-made apparatus to determine both transient- and steady-state values of k from a
single test. Results obtained from three different soils suggest that the steady-state k values are
consistently higher than their transient-state counterparts. For all tested materials, transient-state
k values are sensitive to the rate of soil temperature increment, whereas the values of steady-state
k remain constant irrespective of the rate of transient heating. Numerical simulations of
laboratory experiments suggest that the use of a time-dependent function of k predicts the
experiment results adequately. Moreover, analyses demonstrate that the use of a transient-state k
value may significantly underestimate power output from a geothermal pile.

Keywords: Thermal conductivity, soil, geothermal pile, laboratory test, numerical simulation

INTRODUCTION

Soil thermal conductivity dictates the rate of heat flow in soil. Precise knowledge of soil
thermal conductivity is critical in design and analyses of several buried geoengineering systems
(e.g., borehole heat exchangers, geothermal piles, buried power transmission cables, nuclear
waste repositories). Two common measurement methods for soil thermal conductivity are (a)
transient-state measurement (using a thermal probe or through field thermal response test) that
uses the infinite line heat source model and (b) steady-state measurement (using guarded hot
plate apparatus or divided bar apparatus) that employs Fourier’s law of heat conduction. A
striking disagreement exists in the published literature on the difference between the measured
values of steady-state (kst) and transient-state soil thermal conductivity (ktr).
Nearly identical values of steady- and transient-state soil thermal conductivity are reported
for different types of solid rocks (Sass et al., 1984), rubber compounds (Kerschbaumer et al.,
2019), and dry Tottori sand and Iwami sandy loam (Mahdavi et al., 2019). While Kerschbaumer
et al. (2019) used a guarded heat flow meter for steady-state measurements and a high-pressure
rheometer for transient-state measurements, Mahdavi et al. (2019) employed a spherical steady-
state thermal conductivity measurement apparatus and a single needle thermal probe for transient
measurements. Giordano et al. (2019) noted that thermal conductivity obtained from the transient
line source model could be significantly low as compared to the value of steady-state thermal
conductivity. Abuel-Naga et al. (2009) and Low et al. (2015) postulated some reasons behind the
recorded difference between steady- and transient-state soil thermal conductivity. For soft

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 352

Bangkok clay, Abuel-Naga et al. (2009) reported lower values of kst obtained using a divided-bar
apparatus than ktr values determined using a thermal probe. Such difference in k values was
attributed to the difference in sample size and heat flow pattern. Low et al. (2015) used a thermal
cell and a single needle probe for laboratory thermal conductivity tests on London clay samples
and identified the heat loss from the thermal cell apparatus as the primary reason for high values
of kst. For dry sand, the average values of kst calculated using the steady-state soil thermal
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

response measured in a custom-made laboratory setup are greater than those obtained from the
transient-state portions of the same small-scale thermal response tests; however, the observed
trend was just the opposite for hardened concrete specimen (Kramer et al., 2015).
The reasons behind variations in soil thermal conductivity during transient- and steady-state
measurements may include, but not limited to, the measurement methods, heat flow direction,
power input to (or rate of heating of) the soil sample, and volume of soil specimen tested. The
primary objective of this research is to eliminate the influences of some of these factors on soil
thermal conductivity measurements. The novelty of the series of experiments presented herein
lies in employing a unique thermal conductivity measurement apparatus for determination of
both transient- and steady-state soil thermal conductivity values from a single test. A series of
laboratory thermal conductivity tests were conducted on specimens of saturated Black Cotton
soil and on dry and saturated industrial grade silica sand. The values of ktr and kst obtained from
experiments are further used as input to numerical simulations of laboratory experiments to
identify differences between time-dependent soil temperature profile obtained from experiments
and numerical analyses. Such an exercise reveals the effects of employing a unique value of soil
thermal conductivity, as often done in routine analyses of heat transfer in soil, on soil
temperature response. Furthermore, potential influence of the difference between transient- and
steady-state thermal conductivity values on thermal performance of a geothermal pile is explored
through numerical simulations of pile-soil heat exchange.

MEASUREMENT OF SOIL THERMAL CONDUCTIVITY

Modified consolidometer for soil thermal conductivity measurement

This research uses a stainless-steel consolidation cell (125-mm-diameter and 170-mm-height)


with a centrally placed copper tube. The height, inner and outer diameter of the copper tube are
160 mm, 9.5 mm and 11 mm, respectively. A power regulated heater (with 50W, 230V rating)
can be inserted into the copper tube to heat the soil for thermal conductivity measurement. Soil
temperature changes were measured using thermocouples placed at six pre-decided locations
inside the soil specimen. All six thermocouples were placed on a single horizontal plane. A
screw-clamp arrangement ensured that the thermocouples were placed at their predecided
locations through a pair of nylon strings (Figure 1). Two thermocouples were placed exclusively
on the soil-heater interface (i.e., attached to the surface of the heat source) to imitate needle
probes used for transient thermal conductivity measurements. Figures 1a and 1b show the
schematic representation and the assembled soil thermal conductivity measurement apparatus.
The boundary conditions play crucial role in accurate measurement of both transient- and steady-
state soil thermal conductivity. Thermal insulations on top and bottom of a soil specimen were
achieved, respectively, through the use of a fiberglass (k = 0.04 Wm−1K−1) loading pad and glass
wool (k = 0.03 Wm−1K−1). The dimensions of the mould and the central heating tube follows
ASTM D5334-14 (2014) to avoid the effect of side boundary during transient thermal

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 353

conductivity measurement. The radial boundary was kept as convective boundary to ensure
quick attainment of thermal equilibrium for calculation of kst.
Temperature data were acquired using special T-type thermocouples connected to NI 9213
module at a sampling rate of 1 Hz. The vertical displacement of the sample during consolidation
(for saturated fine-grained soil) or elastic compression (for coarse-grained soil) under the
application of a target vertical load was measured using a dial gauge and a potentiometer
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

connected to NI 9215 module. Both data acquisition modules (i.e., NI 9213 and NI 9215) were
housed in an NI cDAQ 9172 chassis. A digital powermeter was attached to the heater and
voltage regulator for an accurate power display and this arrangement enabled application of a
constant input power to soil samples.

Figure 1. Custom-built thermal conductivity measurement apparatus: (a) schematic


representation (b) assembled consolidation cell with top loading pad

Test materials

In order to study the effect of measurement methods on soil thermal conductivity one fine-
grained and one coarse-grained material were selected for the test program. Black Cotton soil
and dry and saturated industrial grade silica sand were used for thermal conductivity tests. Black
Cotton soil (specific gravity G = 2.65, Liquid limit = 69%, Plasticity index = 26%) is commonly
found in Southern Peninsular region of India and the material for this research is collected from
Malaprabha region (Karnataka) in India. Industrial grade silica sand (specific gravity G = 2.63,
minimum void ratio emin = 0.64, and maximum void ratio emax = 1.0) is poorly graded with sub-
rounded particles and mean particle size equal to 0.5 mm. The thermal conductivity test
apparatus was calibrated using two standard materials, glycerol and fused silica, as mentioned in
ASTM D-5334.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 354

Thermal conductivity tests

For tests reported in this paper, both fine-grained and coarse-grained materials were tested
under a vertical effective stress of 100 kPa. For Black Cotton soil, a slurry was prepared by
mixing dry soil with water in a vacuum aided planetary mixer. The water content of the slurry
was 1.5 times the liquid limit and the slurry was mixed for at least 5 hours for proper saturation.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Thus prepared slurry was poured into the consolidation cell to fill upto an initial height of 150
mm and was kept for self-consolidation for a duration of 24 hours. The self-consolidation phase
was followed by several load increments and subsequent consolidation phases under such load
increments to reach a vertical effective stress level of 100 kPa. For sand, saturation was done by
supplying water from the bottom of the sample compacted within the consolidation mould. The
volume of water passed through the sample was more than five times the volume of voids to
ensure proper saturation of the sample. A vertical load corresponding to achieve vetical effective
stress of 100 kPa on the specimen was applied after saturation. For dry and saturated sand
specimens, relative density was calculated after taking into account of the recorded deformation
caused by the application of the vertical load. On completion of the preparation of a test
specimen under a target vertical stress level, the heater was inserted into the copper tube and a
constant input power was applied to conduct thermal conductivity test (undrained heating) on all
materials.

Calculation of soil thermal conductivity


Most of the past studies focused on transient-state soil thermal conductivity ktr (Abu-Hamdeh
& Reeder, 2000; Chen, 2008; Gangadhara Rao & Singh, 1999; Park et al., 2016; Roshankhah et
al., 2021; Tang et al., 2008; Xiao et al., 2018; Xu et al., 2019; Yu et al., 2016; Zhang et al.,
2018), whereas research aimed at determination of the steady-state value of soil thermal
conductivity kst is rather limited (Abuel-Naga et al., 2009; Alrtimi et al., 2014; Kramer et al.,
2015; Low et al., 2015). The transient-state thermal conductivity ktr is commonly estimated using
the infinite line heat source theory. In the present research, the thermocouples attached to the
surface of the copper tube (heat source) replicate the functionality of a thermal probe. For a heat
input q per unit length of the heater, transient thermal conductivity ktr can be calculated as

q
ktr = (1)
4

where λ is the slope of the straight line segment of the temperature versus natural logarithm of
time profile.
Fourier’s law of heat conduction is used to calculate the steady-state soil thermal conductivity
kst. For the experiment setup used in this research, the heat flow is radially outward from the central
heat source. It took almost 7-8 hours for every sample to reach a non-isothermal equilibrium
condition. To apply Fourier’s law, the top and bottom of the sample was properly insulated,
respectively, using a fiberglass loading pad and glass wool (placed at the bottom of the mould).

r 
Q ln  2 
kst =  r1  (2)
2 L (T1 − T2 )

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 355

where L (m) is the length of the heat source, T1 (K) and T2 (K) are temperature recorded at radial
distances r1 (m) and r2 (m) from the center of the heat source. Figure 2 shows soil temperature
profiles at different locations inside the specimen for a typical thermal conductivity test
performed as part of this research.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Variations of temperature recorded at different thermocouple locations during


thermal conductivity test on saturated Black Cotton soil

Numerical modelling of experiments

The laboratory thermal conductivity tests are simulated using a commercially available finite
element (FE) software COMSOL Multiphysics®. The schematic of the analysis domain and
boundary conditions are shown in Figure 3. The soil domain was modelled as a porous medium
with porosity and height equal to those recorded just before the thermal conductivity tests (i.e., at
the end of consolidation under 100 kPa). The top and bottom boundaries are set as insulated
boundaries, whereas a convective boundary condition is assigned for the radial boundary to
achieve the steady-state condition. Constant-power heat conduction analyses are performed to
obtain soil temperature response with time. Analyses are performed using thermal conductivity
values equal to ktr and kst obtained from the laboratory experiments. A third set of analysis is
performed considering time-dependent variation of soil thermal conductivity from ktr in the
initial phase of heating to kst towards the end of a test.

THERMAL PERFORMANCE ASSESSMENT OF A GEOTHERMAL PILE

Three-dimensional finite element analyses (FEAs) were performed for a double U-tube
geothermal pile to study the effect of soil thermal conductivity on pile thermal performance. The
FE model considers heat transport through three major components of the pile-soil heat exchange

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 356

system – circulation fluid, concrete pile, and saturated soil. A 30×30×25 m3 soil domain with a
centrally placed 450-mm-diameter and 20-m-long geothermal pile is modelled in COMSOL
Multiphysics®. Soil, concrete pile, and circulation tubes (PVC) are modelled as solid domains,
and only conductive heat transport is allowed through these domains. Except for the ground
surface, constant-temperature boundary condition with temperature equal to an initial ground
temperature is assigned for all boundaries. A convective boundary condition is used for the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ground surface. Model geometry, boundary conditions, and schematic plan view of the analysis
domain are shown in Figure 4. Key input parameters used in the analysis are included in Table
1.

nsulated boundary

opper tube
eater
on ecti e boundary
mm

on ecti e boundary oil


1

1 mm

nsulated boundary

Figure 3. Schematic representation of the numerical heat transfer analysis domain

Table 1. Key input parameters for FEAs of a double U-tube geothermal pile

Parameters Value
Initial ground temperature 20 °C
Test duration 30 days
Constant inlet temperature 40 °C
Fluid flow rate 0.15 kgs−1
Thermal conductivity of concrete 1.5 Wm−1K−1
Ground thermal conductivity, kst 3.5 Wm−1K−1
Circulation pipe spacing 150 mm
Shank spacing 75 mm

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 357

on ecti e
Fluid inlet Fluid outlet
boundary

oncrete pile

onstant temperature boundary T


onstant temperature boundary T inlet
ile with

m
embedded

m
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

m
circulation
pipe

.
oil
outlet
. m oil
m
m
onstant temperature boundary T

(a) (b)

Figure 4. Analysis domain for thermal performance assessment of geothermal pile (a)
model geometry and boundary conditions (b) schematic plan view

RESULTS AND DISCUSSIONS


.
Effect of rate of soil temperature increment ΔT on kst and ktr

For all materials tested in this research, the average kst values are greater than ktr (Figure 5a).
The thermal conductivity ratio κ, i.e., the ratio of average steady-state to transient-state values of
soil thermal conductivity, is equal to 1.05, 1.11 and 1.17 for dry sand, saturated Black Cotton
soil, and saturated sand, repectively. An additional set of tests were performed with different
power input to investigate the reason for the observed difference between kst and ktr. Figure 5b
demonstrates. that the values of kst remain unchanged with increase in the rate of .
soil temperature
increment Δ𝑇, whereas the ktr values increase continuously with increase in Δ𝑇 and approach the
value of kst at elevated power input.

(a) (b)

Figure 5. (a) Comparison of ktr and kst for all materials at Q = 20W (b) effect of rate of soil
temperature increment on ktr

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 358

A change in input power alters initial heat flow rate and the thermal response of soil close to
the heater. Thus, the slope λ of temperature versus natural logarithm of time also varies causing a
variation in ktr. For both fine-grained and coarse-grained soils, the thermal conducti ity ratio κ
reduces significantly and approaches the value of unity as the rate of soil temperature change
increases.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Soil thermal conductivity as a function of time

Test results suggest that a unique value of soil thermal conductivity, which is usually an input
to numerical analyses, may not accurately predict soil temperature response. Figure 6 compares
soil temperature responses obtained from numerical analyses and experiments. Analysis using ktr
as an input value of thermal conductivity over-estimates the steady-state soil temperature,
whereas analysis with kst as an input parameter predicts the steady-state temperature reasonably
well but fails to precisely predict soil temperature response prior to the achievement of steady-
state. A simple mathematical function that accounts for a linear transition of k from ktr to kst over
the duration of a test is used for some additional numerical analyses (Eq. 3). Note that such a
variation of k depends on soil type and the rate of soil temperature increment and that the
assumed variation of k may not always be linear.

k tr , 0 to 50 min

 (k - k )
k = k tr + (t - t1 ) st tr , 50 to 200 min (3)
 t 2 - t1
k st , 200 to 450 min

Analysis with k expressed as a function of time predicts recorded soil temperature data with
reasonable accuracy (Figure 6). Such an observation may prove useful for long-term
performance prediction of energy geostructures.

Figure 6. Comparison of soil temperature response recorded during thermal conductivity


test on Black Cotton soil and those obtained from numerical analysis

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 359

Effect of soil thermal conductivity on geothermal pile performance

Analysis of data obtained from thermal conductivity tests concluded that the steady-state soil
thermal conductivity is greater than that at transient-state. The effect of such a variation in soil
thermal conductivity during the operational duration of a geothermal pile with double U-tube on
power output is explored herein. For different  values, Figure 7 shows the variation of power
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

output over an operational period of 30 days. For a soil with higher value of , consideration of
ktr as an input parameter for numerical performance analysis will under predict power output of a
geothermal pile. While the use of ktr can be ideal for short-term performance assessment, the use
of kst would be more appropriate for long-term performance assessment for which soil near the
geothermal pile is likely to reach a quasi-steady state.

Figure 7. Variation of power output of a single geothermal pile for different  values

CONCLUSIONS

Results from thermal conductivity tests and numerical analyses presented in this paper
identify the importance of considering time-dependent variations of soil thermal conductivity for
applications involving continuous heat transfer in ground. Transient- and steady-state soil
thermal conductivity values are obtained, under the same test boundary conditions, using a
specially designed thermal conductivity measurement device. Data gathered from experiments
on three different soils suggest that the steady-state soil thermal conductivity values are
consistently higher than the transient-state values. For all materials tested in this research, the
values of steady-state thermal
.
conductivity remain constant irrespective of the rate of soil
temperature increment Δ𝑇.
during transient heating, whereas the transient-state k values increase
with an increase in Δ𝑇 and approach the steady-state value. Numerical simulations of the
laboratory experiments suggest that the use of a time-dependent function of soil-thermal
conductivity may predict soil temperature response more accurately than that predicted using a
unique value of soil thermal conductvty. Furthermore, results demonstrate that a uniform value

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 360

of k, as often used in thermal performance assessment of geothermal piles, may significantly


underestimate power output under continuous operation.

REFERENCES

Abu-Hamdeh, N. H., and Reeder, R. C. (2000). Soil thermal conductivity effects of density,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

moisture, salt concentration, and organic matter. Soil science society of America Journal,
64(4), 1285-1290.
Abuel-Naga, H. M., Bergado, D. T., Bouazza, A., and Pender, M. J. (2009). Thermal
conductivity of soft Bangkok clay from laboratory and field measurements. Engineering
Geology, 105(3-4), 211-219.
Alrtimi, A., Rouainia, M., and Manning, D. (2014). An improved steady-state apparatus for
measuring thermal conductivity of soils. International Journal of Heat and Mass Transfer,
72, 630-636.
ASTM. ASTM D5334-14. (2014). D5334-14 Standard Test Method for Determination of
Thermal Conductivity of Soil and Soft Rock by Thermal Needle Probe Procedure. ASTM
International: West Conshohocken, PA, USA.
Chen, S. X. (2008). Thermal conductivity of sands. Heat and mass transfer, 44(10), 1241-1246.
Gangadhara Rao, M., and Singh, D. (1999). A generalized relationship to estimate thermal
resistivity of soils. Canadian Geotechnical Journal, 36(4), 767-773.
Ghasemi-Fare, O., and Basu, P. (2018). Influences of ground saturation and thermal boundary
condition on energy harvesting using geothermal piles. Energy and Buildings, 165, 340-351.
Giordano, N., Chicco, J., Mandrone, G., Verdoya, M., and Wheeler, W. H. (2019). Comparing
transient and steady-state methods for the thermal conductivity characterization of a borehole
heat exchanger field in Bergen, Norway. Environmental Earth Sciences, 78(15), 1-15.
Kerschbaumer, R. C., Stieger, S., Gschwandl, M., Hutterer, T., Fasching, M., Lechner, B.,
Meinhart, L., Hildenbrandt, J., Schrittesser, B., and Fuchs, P. F. (2019). Comparison of
steady-state and transient thermal conductivity testing methods using different industrial
rubber compounds. Polymer testing, 80, 106121.
Kramer, C. A., Ghasemi-Fare, O., and Basu, P. (2015). Laboratory thermal performance tests on
a model heat exchanger pile in sand. Geotechnical and Geological Engineering, 33(2), 253-
271.
Low, J. E., Loveridge, F. A., Powrie, W., and Nicholson, D. (2015). A comparison of laboratory
and in situ methods to determine soil thermal conductivity for energy foundations and other
ground heat exchanger applications. Acta geotechnica, 10(2), 209-218.
Mahdavi, S., Neyshabouri, M., and Fujimaki, H. (2019). A Spherical Steady-State Method to
Measure Soil Thermal Conductivity. Eurasian Soil Science, 52(12), 1572-1576.
Park, K., Lee, J., Yoon, H.-K., and Kim, D. (2016). Hydraulic and thermal conductivities of
kaolin–silica mixtures under different consolidation stresses. Marine Georesources &
Geotechnology, 34(6), 532-541.
Roshankhah, S., Garcia, A. V., and Carlos Santamarina, J. (2021). Thermal Conductivity of
Sand–Silt Mixtures. Journal of Geotechnical and Geoenvironmental Engineering, 147(2),
06020031.
Sass, J., Stone, C., and Munroe, R. J. (1984). Thermal conductivity determinations on solid
rock—a comparison between a steady-state divided-bar apparatus and a commercial transient
line-source device. Journal of Volcanology and Geothermal Research, 20(1-2), 145-153.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 361

Tang, A.-M., Cui, Y.-J., and Le, T.-T. (2008). A study on the thermal conductivity of compacted
bentonites. Applied Clay Science, 41(3-4), 181-189.
Xiao, Y., Liu, H., Nan, B., and McCartney, J. S. (2018). Gradation-dependent thermal
conductivity of sands. Journal of Geotechnical and Geoenvironmental Engineering, 144(9).
Xu, Y., Zeng, Z., and Lv, H. (2019). Effect of temperature on thermal conductivity of lateritic
clays over a wide temperature range. International Journal of Heat and Mass Transfer, 138,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

562-570.
Yu, X., Zhang, N., Pradhan, A., and Puppala, A. J. (2016). Thermal conductivity of sand–kaolin
clay mixtures. Environmental Geotechnics, 3(4), 190-202.
Zhang, M., Lu, J., Lai, Y., and Zhang, X. (2018). Variation of the thermal conductivity of a silty
clay during a freezing-thawing process. International Journal of Heat and Mass Transfer,
124, 1059-1067.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 362

Exploring Box Fixity and Platen Texture in Large-Scale Direct Shear Testing

Nicholas A. Culbreth1; Jennifer E. Nicks, Ph.D., P.E., M.ASCE2; Michael T. Adams, M.ASCE3;
and Thomas Gebrenegus, Ph.D.4
1
High Performance Technologies, Inc., Herndon, VA. Email: nicholas.culbreth.ctr@dot.gov
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Federal Highway Administration, Turner-Fairbank Highway Research Center, McLean, VA.
Email: jennifer.nicks@dot.gov
3
Federal Highway Administration, Turner-Fairbank Highway Research Center, McLean, VA.
Email: mike.adams@dot.gov
4
High Performance Technologies, Inc., Herndon, VA. Email: t.gebrenegus.ctr@dot.gov

ABSTRACT

The Federal Highway Administration (FHWA) recently completed an interlaboratory round-


robin study that explored the shear strength variability of open-graded aggregates (OGAs) in
large-scale direct shear (LSDS) devices. The study found that many participating labs had
significantly different device configurations and test procedures. Unlike standard direct shear
devices (box width ≤100 mm), there are no unique industry standards for LSDS device design
(box width ≥300 mm) or test procedures. Thus, LSDS testing remains the Wild West compared
to more established standard direct shear test practices. To explore this frontier, FHWA’s
geotechnical laboratory is experimentally and numerically investigating key variables of LSDS
device design. This paper presents results from FHWA’s experimental investigation into the
influence of upper shear box fixity and the texture of a hinged load platen on the deformation
behavior and strength of a selected OGA. Results indicate that box fixity has a significant
influence, with mobile boxes producing peak shear stresses approximately half those of fixed
boxes. In contrast, platen texture had negligible influence. Ultimately, this paper seeks to begin
an in-depth discussion of the questions prompted by the round-robin study—which device
configurations produce the most consistent results, and which configurations best represent an
aggregate’s true behavior and strength?

INTRODUCTION

The strength of granular soils and aggregates is often an integral component required for the
design phases of transportation projects. Determining the strength of a granular material is
critically important when calculating lateral earth pressures, bearing resistance, settlement, and
other strength or service limit states. Direct shear tests offer a simple, fast, and nominally
repeatable method for measuring the strength of granular materials. Given the geotechnical
significance of accurately evaluating a material’s strength, it is no surprise that standard direct
shear (SDS) testing of sands and other fine-grained materials is performed in accordance with
well-established, industrywide standards. These standards include ASTM D3080-11 (2012) and
AASHTO T 236-08 (2018). However, such industry standards have seemingly been adopted for
LSDS testing of coarser materials (e.g., open-graded aggregates (OGAs), well-graded crushed
stone, road bases, ballast) without consideration for the differences in behavior between fine- and
coarse-grained materials. LSDS device manufacturers and testing laboratories have therefore
adapted SDS device schematics and procedures to large-scale applications with varying
modifications.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 363

Shibuya et al. (1997) previously categorized the variety in SDS device designs into three
primary design archetypes (see Figure 1), with the largest differences related to the mobility of
the upper shear box, the mobility of the load platen, and the fixity of the connection between the
normal load source and load platen. Recent FHWA research through an LSDS round-robin study
(Nicks et al. 2022) showed a wider diversity of shear boxes in use; only two of the six
laboratories involved in the round-robin fit within the existing SDS classifications described by
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Shibuya et al. (1997). The remaining four boxes were hybrid designs, implementing various
features of the different box archetypes. Additional design choices were noted through the round-
robin study—such as asymmetric shear box halves; smooth or textured sidewalls, box floors, and
load platen surfaces; direction of the shear force (i.e., pushing or pulling); and location of the
shear force (i.e., the upper or lower box).

Source: FHWA.

Figure 1. Basic types of direct shear boxes, as described by Shibuya et al. in 1997.

Industrywide inconsistency in LSDS device design is mirrored in the test procedures applied.
Testing laboratories are left to adapt their own procedures or follow manufacturer
recommendations; both instances often reflect a scaled-up version of SDS test methods. Nicks et
al. (2022) found that test procedures varied in much the same manner as device design, with
similar deviations from SDS procedures. Examples included differing sample aspect ratios (e.g.,
2:1 width to height, per ASTM D3080-11 (2012)), compaction methods (hand tamping or
vibratory table), and shear gap timings (i.e., pre- or post-consolidation).
Varying LSDS device designs and test procedures ultimately limit the interlaboratory
reproducibility of results and lead to variability in the deformation behavior of aggregates during
shear and the resulting shear strength parameters (Nicks et al. 2022). Notably, differences in
tangent friction angles (ϕt) ranged from 25 to 47 percent between the six laboratories
participating in the round-robin. The interlaboratory variability in LSDS test results found by
Nicks et al. (2022) is contrasted by, relatively speaking, much lower variability in measured
intrinsic material properties, such as differences between 3.3 and 9.6 percent for measured dry
unit weights at 95 percent relative density.
Prior research on SDS test variability further highlights the need to reign in LSDS testing.
Bareither et al. (2008) conducted a similar interlaboratory variability study on SDS tests of
sands, finding variability comparable to the LSDS results reported by Nicks et al. (2022), with
percent differences between 30.9 and 54.2 for ϕt. In contrast, intralaboratory results from
Bareither et al. (2008) found variability to be minimal and statistically insignificant. Similarly,
Thermann et al. (2006) found no statistically significant difference in results from two separate
test operators conducting SDS tests using the same device and procedure. These findings suggest

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 364

that different test operators may not be a main contributing factor to the variability found by
Nicks et al. (2022) for LSDS devices; however, further investigation into LSDS operator
variability is necessary. Ultimately, Nicks et al. (2022) found that interlaboratory variability was
significantly impacted by LSDS device characteristics—particularly upper shear box fixity,
which was strongly correlated with higher strength results.
To that end, FHWA initiated a new study with the objective of standardizing LSDS testing,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

including making recommendations for a practical LSDS device configuration that produces
consistent and representative results for coarse materials. Within this study, a series of tests were
performed on the existing LSDS device in the FHWA Geotechnical Laboratory using a single
test operator, specimen, and test procedure with either fixed or mobile upper shear boxes and a
smooth or textured load platen surface. This paper is the first exploration of this study that
directly investigates the influence of two selected LSDS device configuration variables on shear
strength results. It also explores future research needs in broader LSDS practice refinement.

TEST PROGRAM

Material. A No. 8 aggregate per the AASHTO M 43-05 (2018) specifications with siltstone
(SI) mineralogy was procured from the same quarry source as the No. 8-SI material tested in the
Nicks et al. (2022) study. This sample was selected for the consistently low interlaboratory
variability of its intrinsic material properties (e.g., maximum dry density) yet higher variability in
shear strength results (e.g., tangent friction angle). Immediately after the bulk 5-ton delivery of
aggregate, buckets were filled from three sides of the stockpile, collected, and labeled. Equal
numbers of buckets from each side of the pile were then oven dried for 24 hours before blending
and quartering per AASHTO’s T 248-14 (2014) standard. The blended and quartered samples
were then used for the sieve analysis (ASTM C136-19 (2019)), for minimum and maximum unit
weight (ASTM D4254-16 (2016) and D4253-16 (2016), respectively), and for large-scale direct
shear tests (ASTM D3080-11 (2012)). The minimum, maximum, and dry unit weights at 95
percent relative density were 13.4, 15.7, and 15.6 kN/m3, respectively. Sieve analysis found that
50 percent (D50) of material was finer than 7.6 mm, and 85 percent (D85) of material was finer
than 10.7 mm. As shown in Figure 2, the sample material procured did not meet AASHTO M
43-05 (2018) limits for a No. 8 aggregate, but testing proceeded, as this study’s focus was on
broader LSDS coarse material testing and was not specific to No. 8 aggregates or SI mineralogy.

Source: FHWA.

Figure 2. No. 8-SI gradation results.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 365

The standard setup of the FHWA Geotechnical Laboratory’s LSDS device is a fixed upper
shear box (FB) and smooth platen (SP) configuration. Modifications were necessary to facilitate
mobile upper box (MB) and grooved platen (GP) configurations. The mobile upper box
configurations were achieved by removing the braces that restrain vertical displacement of the
upper shear box during testing (denoted by red hashmarks in Figure 3). Additional linear variable
differential transformers (LVDTs) were mounted to the LSDS device to measure any change in
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

vertical displacement of the upper shear box during testing and to verify mobility.

Source: FHWA.
Green hashmarks indicate fixity; red hashmarks indicate removable fixity.

Figure 3. Diagram of FB-SP configuration for LSDS device.

The grooved platen had to be machined by FHWA; the design was based on discrete element
method modeling of soil-structure interfaces from Jing et al. (2017). Their modeling found that
shear strength increases as roughness of the structure (i.e., load platen) increases, with
diminishing returns once the wavelength of a 45-degree sawtooth structure reaches the median
grain size (i.e., D50) of the test material. Accordingly, FHWA manufactured a grooved load
platen with a 45-degree sawtooth texture and a wavelength approximately equal to the average
D50 of all No. 8 aggregates from Nicks et al. (2022), which corresponded to the same D50 as the
No. 8-SI selected for this study (7.6 mm).
Test Procedures. Three LSDS test series (labeled S1, S2, and S3) were conducted for each
device configuration: fixed box with grooved platen (FB-GP), fixed box with smooth platen (FB-
SP), mobile box with grooved platen (MB-GP), and mobile box with smooth platen (MB-SP).
Each test series consisted of individual tests performed at target normal stresses of 34.5, 103.4,
and 206.8 kPa. A total of 9 tests were conducted per configuration, with 36 tests completed for
the entire program.
Samples were prepared in three 50-mm lifts of equal mass for a total thickness of 150 mm,
meeting the ASTM D3080-11 (2012) 2:1 aspect ratio requirement; a 25.4-mm spacer was placed
at the bottom of the box to ensure equal sample height above and below the shear plane. Between
each lift, the samples were compacted to a dry unit weight at 95 percent relative density (15.6
kN/m3) by hand tamping. Tests were performed under dry conditions. Samples were sheared at a
rate of 0.38 mm per minute until termination at 20 percent horizontal strain (equivalent to 61

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 366

mm). The shear gap was set to the samples’ D85 (10.7 mm) after the consolidation phase.
Vertical LVDT measurements were used to verify that the shear gap was within 0.25 mm of the
target gap value. Box displacement during the consolidation phase was minimal for all LSDS
device configurations and was assumed to be zero for gap verification purposes.
A consolidation phase time of 5 min was established, with each test achieving 90 percent
consolidation (t90) within 5 min. Data was collected every second during the consolidation phase
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

and every 10 s during the shear phase. The peak shear stress was defined as the maximum
nominal shear stress achieved during a test. Tangent friction angles were calculated from the
slope of the linear Mohr-Coulomb failure envelopes using the least-squares regression method.
Positive displacement was defined as sample contraction, with negative displacement equaling
dilation. Target normal stresses were used in place of the measured nominal normal stresses for
all calculations. No area, displacement, or device resistance corrections were performed. Tilt
angles for the upper shear box and the load platen were calculated as the arctangent of the
difference in average displacement between the resistance and load sides divided by the
longitudinal distance between the relevant LVDTs (see Figure 3). Displacement, and by
extension tilt, was zeroed to the initial value at the start of the shear phase as displacement (and
tilting) were negligible during consolidation for all device configurations.

RESULTS AND DISCUSSION

Shear Strength. To illustrate the results, stress-strain curves for tests conducted at the target
normal stress of 103.4 kPa are presented in Figure 4-a.

Source: FHWA.

Figure 4. (a) Stress-strain curves under 103.4 kPa normal stress, (b) Averaged Mohr-
Coulomb failure envelopes.

Table 1 lists the peak shear stresses and corresponding horizontal strain values alongside
basic summary statistics of mean (x̄), standard deviation (s), and coefficient of variation (COV)
for all test series (Sn). Peak shear stresses for FB configurations were up to double those of their
MB counterparts and occurred later in the shear phase.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 367

Table 1. Peak shear stress and horizontal strain results.

Target Horizontal Strain at Peak Shear


Peak Shear Stress (kPa)
Normal Test (percent)
Stress Number MB- MB- MB-
FB-GP FB-SP MB-SP FB-GP FB-SP
(kPa) GP GP SP
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

S1 118.0 119.0 56.3 56.7 6.2 5.9 3.6 5.6


S2 148.1 144.2 59.5 55.7 5.6 5.9 4.9 4.6
S3 139.8 132.9 57.1 57.4 6.0 7.1 3.9 5.4
34.5
x̄ 135.3 132.0 57.6 56.6 5.9 6.3 4.1 5.2
s 12.7 10.3 1.4 0.7 0.3 0.6 0.6 0.4
COV 9.4% 7.8% 2.4% 1.2% 5.1% 9.5% 14.6% 7.7%
S1 269.0 241.3 144.4 152.5 6.4 6.3 5.1 5.6
S2 274.4 310.2 139.7 153.3 6.5 7.3 5.6 6.3
S3 261.5 253.4 141.3 153.8 7.8 9.6 6.1 5.6
103.4
x̄ 268.3 268.3 141.8 153.2 6.9 7.7 5.6 5.8
s 5.3 30.0 2.0 0.5 0.6 1.4 0.4 0.3
COV 2.0% 11.2% 1.4% 0.3% 8.7% 18.2% 7.1% 5.2%
S1 378.3 451.6 262.7 275.1 7.6 8.8 6.1 6.3
S2 436.8 382.4 258.4 272.9 8.2 8.8 6.5 6.3
S3 407.4 447.4 265.8 265.1 7.5 8.4 7.8 6.8
206.8
x̄ 407.5 427.1 262.3 271.0 7.8 8.7 6.8 6.5
s 23.9 31.7 3.0 4.3 0.3 0.2 0.7 0.2
COV 5.9% 7.4% 1.1% 1.6% 3.8% 2.3% 10.3% 3.1%

Figure 4-b illustrates the linear Mohr-Coulomb failure envelopes averaged from all test series
for each configuration. Tangent friction angles and corresponding apparent cohesion values from
individual test series can be found in Table 2. Tangent friction angles from the FB configurations
averaged a 7- to 9-degree increase over the MB setups, comparable to the findings of Nicks et al.
(2022). Differences between the GP and SP configurations for all strength parameters were
significantly smaller than those between the FB and MB configurations. Strength parameter
COVs were lowest among the MB configurations—e.g., ϕt COVs averaged 1.2 percent for
mobile boxes, as opposed to 4.8 percent for fixed. The low COVs suggest the MB setups offered
more repeatable results.

Table 2. Tangent friction angle and apparent cohesion results.

Test Tangent Friction Angle (degree) Apparent Cohesion (kPa)


Number FB-GP FB-SP MB-GP MB-SP FB-GP FB-SP MB-GP MB-SP
S1 55.9 62.7 50.0 51.6 85.6 48.0 17.3 16.6
S2 59.0 53.0 49.1 51.4 94.9 126.4 20.0 16.8
S3 57.0 61.3 50.4 50.1 92.4 67.7 15.6 21.4
x̄ 57.3 59.0 49.8 51.0 91.0 80.7 17.6 18.3
s 1.3 4.3 0.5 0.7 3.9 33.3 1.8 2.2
COV 2.3% 7.3% 1.0% 1.4% 4.3% 41.3% 10.2% 12.0%

Displacement. Platen displacement (see Figure 5-a) during the first half of shear (i.e., <10
percent horizontal strain) was similar across all configurations.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 368
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Source: FHWA.

Figure 5. (a) Platen and (b) Box vertical displacement under 103.4 kPa normal stress.

However, the mobile upper box setups reached a lower maximum platen displacement than
their fixed upper box counterparts. As expected, box displacement (see Figure 5-b) was minimal
for the FB configurations. Meanwhile, MB tests saw box displacement match or exceed that of
platen displacement for every test (see Table 3).

Table 3. Maximum vertical displacement during shear.

Target Maximum Platen Displacement


Maximum Box Displacement (mm)
Normal Test (mm)
Stress Number FB- MB- MB-
FB-SP MB-SP FB-GP FB-SP MB-SP
(kPa) GP GP GP
S1 -14.8 -13.0 -10.8 -10.7 -0.1 -0.1 -12.1 -10.7
S2 -14.8 -15.3 -10.9 -10.0 -0.2 -0.2 -13.6 -10.0
S3 -15.4 -14.4 -9.7 -10.1 -0.2 -0.1 -10.3 -11.4
34.5
x̄ -15.0 -14.2 -10.5 -10.3 -0.2 -0.1 -12.0 -10.7
s 0.3 0.9 0.5 0.3 0.1 0.1 1.3 0.6
COV 2.0% 6.3% 4.8% 2.9% 50.0% — 10.8% 5.6%
S1 -11.1 -10.6 -8.6 -8.2 -0.4 -0.3 -12.1 -10.3
S2 -11.4 -12.0 -8.9 -8.4 -0.2 -0.4 -12.5 -10.5
S3 -10.5 -10.9 -8.9 -8.7 -0.4 -0.3 -11.2 -12.0
103.4
x̄ -11.0 -11.2 -8.8 -8.4 -0.3 -0.3 -11.9 -10.9
s 0.4 0.6 0.1 0.2 0.1 0.0 0.5 0.8
COV 3.6% 5.4% 1.1% 2.4% 33.3% 0.0% 4.2% 7.3%
S1 -8.0 -8.6 -7.1 -6.3 -0.6 -0.6 -9.7 -8.9
S2 -7.8 -6.9 -7.1 -6.7 -0.7 -0.5 -9.1 -9.3
S3 -7.5 -8.4 -7.3 -7.4 -0.8 -0.6 -10.2 -9.7
206.8
x̄ -7.8 -8.0 -7.2 -6.8 -0.7 -0.6 -9.7 -9.3
s 0.2 0.8 0.1 0.5 0.1 0.1 0.5 0.3
COV 2.6% 10.0% 1.4% 7.4% 14.3% 16.7% 5.2% 3.2%
—COV incalculable due to x̄ = 0.00.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 369

The slight displacement observed during FB tests ceased around 6 to 8 percent horizontal
strain, coinciding with peak shear. Mobile boxes saw a distinct contraction at the start of the test
as they displaced into the initial shear gap. The total box displacement observed in the mobile
setup was approximately equivalent to the initial shear gap of 10.7 mm, resulting in a doubling of
the shear plane thickness over the course of the test. These results may explain the reduction in
peak shear stresses compared to the fixed setup. No significant differences were observed in
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

platen or box displacements between grooved and smooth platens, suggesting platen texture
plays a minimal role in deformation behavior.
Shibuya et al. (1997) suggested that friction between the sample and interior box walls is
greatest in fixed box LSDS device designs, ultimately leading to higher-strength measurements.
This hypothesis is corroborated by the tandem displacement of the mobile upper shear box and
platen throughout the peak phase of shear, which would mitigate any resistive force from internal
wall friction. Prior SDS research on poorly graded, fine-to-medium sands using fixed box setups
also found a reduction in shear strength associated with increasing shear gap size due to material
loss through the gap (Lings and Dietz 2004; Kim et al. 2012). Greater material loss for the MB
configurations was observed during testing; however, the impact of material loss in large-scale
testing, particularly for OGAs, needs further study. Geometric changes of the shear zone with a
continually increasing gap for the mobile box provide another potential explanation.
Tilting. The longitudinal tilt of both the platen and the box for each device configuration was
calculated based on the LVDT measurements and their respective locations (see Table 4).

Table 4. Maximum tilting during shear.

Target Maximum Platen Longitude Tilt Maximum Box Longitude Tilt


Norma Test (degree) (degree)
l Stress Number MB- MB- MB- MB-
(kPa) FB-GP FB-SP FB-GP FB-SP
GP SP GP SP
S1 4.72 5.64 4.52 6.12 0.05 0.07 0.91 0.81
S2 5.47 5.00 5.25 5.11 0.09 0.08 0.70 0.59
34.5 S3 5.47 5.01 4.21 5.15 0.07 0.07 0.64 0.84
x̄ 5.22 5.22 4.66 5.46 0.07 0.07 0.75 0.75
s 0.35 0.30 0.44 0.47 0.02 0.01 0.12 0.11
COV 6.7% 5.8% 9.4% 8.6% 28.6% 14.3% 16.0% 14.7%
S1 5.27 5.64 5.64 5.31 0.13 0.10 0.98 0.86
S2 5.57 5.83 5.65 5.74 0.12 0.18 1.03 1.03
S3 4.94 6.57 5.43 5.93 0.11 0.12 1.02 1.10
103.4
x̄ 5.26 6.01 5.57 5.66 0.12 0.13 1.01 1.00
s 0.26 0.40 0.10 0.26 0.01 0.03 0.02 0.10
COV 4.9% 6.7% 1.8% 4.6% 8.3% 23.1% 2.0% 10.0%
S1 6.52 6.90 5.25 5.97 0.21 0.24 1.27 1.05
S2 6.56 7.04 5.60 6.34 0.25 0.27 1.12 1.06
S3 6.46 6.62 5.60 6.05 0.24 0.24 1.12 1.12
206.8
x̄ 6.51 6.85 5.48 6.12 0.23 0.25 1.17 1.08
s 0.04 0.17 0.17 0.16 0.02 0.01 0.07 0.03
COV 0.6% 2.5% 3.1% 2.6% 8.7% 4.0% 6.0% 2.8%

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 370

Platen longitudinal tilt continuously increased throughout shear at a similar rate and extent
for all device configurations (see Table 4 and Figure 6-a); tilting was expected, considering the
platen to normal load connection was hinged.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Source: FHWA.

Figure 6. (a) Platen tilt and (b) Box longitudinal tilt under 103.4 kPa normal stress.

In contrast, box longitudinal tilt varied dramatically between FB and MB tests (see Table 4
and Figure 6-b). The FB configurations experienced minimal box tilting until peak shear was
reached, after which tilting ceased. This trend matched box displacement results from FB tests
(see Figure 5-b).
More interestingly, the MB configurations saw a sharp peak failure in box tilt, which
occurred at the peak shear stress of the No. 8-SI, suggesting a relaxation of box-resistive forces
with movement of the mobile upper box as an explanation of the lower shear strength produced
by MB configurations.

CONCLUSIONS

Current LSDS testing practices vary widely in the absence of specific industry standards.
Nicks et al. (2022) found significant interlaboratory variability in LSDS device design, test
procedure, and strength and deformation behavior. The development of widely accepted LSDS
standards offers a compelling and obvious path to improving LSDS reproducibility. Through a
preliminary experimental study investigating the impact of upper shear box mobility and the
smoothness of the loading platen, the following conclusions were reached:
• The mobile upper shear box configurations tested produced more consistent results but at
a lower magnitude than the fixed upper shear box configurations. Potential explanations
for the reduced strength results of the MB setups included lower resistance to internal box
wall friction, titling of the shear box and platen, and geometric changes to the shear gap.
• Platen texture did not significantly influence strength or deformation behavior in any
test.
• MB design created a dynamic shear plane geometry despite nominally improved
repeatability, potentially sacrificing accuracy for precision.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 371

These first explorational steps have highlighted the great effort still needed to improve and
standardize LSDS testing. FHWA plans further testing of FB and MB configurations, along with
numerical modeling, to evaluate what configuration produces results that most accurately reflect
a granular material’s true strength. These tests may include additional load cells to measure shear
load on the resistance side of the upper shear box and normal load at the base of the sample,
tactile pressure sensors to evaluate pressure distribution, and additional LVDTs to track lateral
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

displacement of the upper shear box. Future testing will also involve investigating additional
LSDS device configuration variables (e.g., hinged versus fixed platen connections). FHWA will
continue its mission to tame the Wild West of LSDS testing.

REFERENCES

AASHTO. (2018). M 43-05: Standard Specification for Sizes of Aggregate for Road and Bridge
Construction, AASHTO, Washington, DC.
AASHTO. (2018). T 236-08: Standard Method of Test for Direct Shear Test of Soils under
Consolidated Drained Conditions, AASHTO, Washington, DC.
AASHTO. (2014). T 248-14: Standard Method of Test for Reducing Samples of Aggregate to
Testing Size. AASHTO, Washington, DC.
ASTM. (2019). ASTM C136-19: Standard Test Method for Sieve Analysis of Fine and Coarse
Aggregates. ASTM, West Conshohocken, PA.
ASTM. (2016). ASTM D4253-16: Standard Test Methods for Maximum Index Density and Unit
Weight of Soils Using a Vibratory Table. ASTM, West Conshohocken, PA.
ASTM. (2016). ASTM D4254-16: Standard Test Methods for Minimum Index Density and
Unit Weight of Soils and Calculation of Relative Density. ASTM, West Conshohocken,
PA.
ASTM. (2012). ASTM D3080-11: Standard Test Method for Direct Shear Test of Soils Under
Consolidated Drained Conditions. ASTM, West Conshohocken, PA.
Bareither, C. A., Benson, C. H., and Edil, T. B. (2008). “Reproducibility of Direct Shear Tests
Conducted on Granular Backfill Materials.” ASTM Geotechnical Testing Journal, 31(1), 84-
94.
Jing, X., Zhou, W., and Li, Y. (2017). “Interface Direct Shearing Behavior Between Soil and
Saw-Tooth Surfaces by DEM Simulation.” Procedia Engineering, 175, 36-42. X, X, X.
Hack, R. R. G. K. (2018). “Mohr-Coulomb Failure Envelope.” Encyclopedia of Engineering
Geology, Encyclopedia of Earth Sciences Series, Bobrowsky P.T., and Marker, B., eds,
Springer, Cham, Capital Region of Denmark.
Kim, B., Shibuya, S., Park, S., and Kato, S. (2012). “Effect of Opening on the Shear Behavior of
Granular Materials in Direct Shear Test.” KSCE Journal of Civil Engineering, 16(17), 1132-
1142.
Lings, M. L., and Dietz, M. S. (2004). “An Improved Direct Shear Apparatus for Sand.”
Geotechnique, 54(4), 245-256.
Nicks, J. E., Adams, M. T., Culbreth, N. A., and Gebrenegus, T. (Forthcoming). Variability of
Engineering Properties of Open-Graded Aggregate Backfills Through Large-Scale Direct
Shear Testing: A Round-Robin Study. FHWA, Washington, DC.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 372

Shibuya, S., Mitachi, T., and Tamate, S. (1997). “Interpretation of Direct Shear Box Testing of
Sands as Quasi-Simple Shear.” Geotechnique, 47(4), 769-790.
Thermann, K., Gau, C., and Tiedemann, J. (2006). “Shear Strength Parameters From Direct
Shear Tests—Influencing Factors and Their Significance.” 10th Congress of the
International Association for Engineering Geology and Environment, paper 484.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 373

Examination of Cone Penetration in Non-Plastic Silt with a Direct Cone Penetration Model

Diane M. Moug, Ph.D.1; and Adam B. Price, Ph.D., P.E.2


1
Dept. of Civil and Environmental Engineering, Portland State Univ., Portland, OR.
Email: dmoug@pdx.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Fugro, Walnut Creek, CA. Email: a.price@fugro.com

ABSTRACT

The cone penetration test (CPT) is widely used to characterize the behavior and properties of
soils ranging from clays to sands. Although the CPT is often used to characterize silts, there exist
few methods to interpret CPT data that were specifically developed for silts. This is partly
attributed to a limited number of numerical and experimental cone penetration studies that
specifically focus on silts. Past studies generally examined either undrained penetration in near-
normally consolidated clay or drained penetration in clean sand, whereas cone penetration in
silty soils (at the standard penetration rate of 2 cm/s) may be drained, partially drained, or
undrained, depending on the soil’s permeability and compressibility. Consequently, uncertainty
is introduced into CPT data interpretation for intermediate soils since geotechnical engineering
practice lacks an established theoretical basis for evaluating cone penetration data in silts and
other intermediate soils such as silty or clayey sands. This study takes a step towards establishing
a theoretical basis by examining cone penetration in a non-plastic silt with a direct axisymmetric
penetration model and the MIT-S1 constitutive model, calibrated against laboratory element
testing and geotechnical centrifuge model tests. The objectives of this study are to (1) validate
the numerical model for non-plastic silt with existing experimental cone penetration data from
centrifuge tests and (2) use the numerical model to examine how simulated CPT data are affected
by drained or undrained conditions during penetration, state-dilatancy relationships, and initial
soil state.

INTRODUCTION

The cone penetration test (CPT) is widely used in geotechnical engineering to characterize
stratigraphy, soil behavior type, and engineering properties. Methods for interpreting CPT data
have largely been developed from synthesis of data from CPTs and laboratory tests, as well as a
variety of other in-situ measurements and observations from case studies, (Boulanger & Idriss
2016, Robertson 2016, Saye et al. 2021), theoretical studies of cone penetration (Senneset et al.
1989, Chen & Mayne 1994), and experimental studies (Houlsby & Hitchman 1988, DeJong &
Randolph 2012). These approaches primarily focused on developing relationships for
sedimentary clays and undrained cone penetration conditions, or for clean sands and drained
penetration conditions. There are far fewer established methods and data sets for soils
intermediate to sands and clays, including non-plastic silts. Consequently, the current standard of
practice for intermediate soils is to use CPT data to estimate whether the soil behavior is
consistent with sand or clay and then proceed with interpretation methods developed for either
undrained penetration in clays or drained penetration in sands.
Interpretation methods for CPT data that were developed for undrained penetration in clays
or drained penetration in sands may not be appropriate for non-plastic silts and intermediate soils

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 374

for several reasons including: (1) penetration conditions may not be entirely undrained or drained
at the standard penetration rate of 2 cm/s (Krage & DeJong 2016; Schnaid et al 2010), and
(2) the compression and shear behaviors of non-plastic silts may not be well represented by the
data sets or assumptions used to develop interpretation methods for sands and clays. Further
study of cone penetration in non-plastic silt is needed to develop robust interpretation methods
and evaluate the reasonableness of the current standard of practice.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

This study examines cone penetration in non-plastic silt with cone penetration simulations,
laboratory test data, and geotechnical centrifuge modeling with in-flight cone penetration testing.
The objectives of this study are to (1) validate a numerical model for non-plastic silt with
existing experimental cone penetration data from centrifuge tests, and (2) examine how
simulated CPT cone penetration resistance (qt) is affected by undrained or drained conditions
during penetration, state-dilatancy relationships and initial soil state. Cone penetration
simulations for non-plastic silica silt are performed using an axisymmetric direct cone
penetration model and the MIT-S1 constitutive model. Three MIT-S1 calibrations are developed
to capture behaviors known to affect cone penetration tip resistance. The calibrations are based
on laboratory shear and compression test data and well-established empirical relationships; they
differ in their prioritization of key behaviors and the available test data. Undrained and drained
penetration is simulated from different initial soil states for each calibration. In-flight cone
penetration test data from a geotechnical centrifuge model are used to validate the cone
penetration model in non-plastic silt.

EXPERIMENTAL CHARACTERIZATION OF A NON-PLASTIC SILT

Soil tested. The soil used in this study was a manufactured non-plastic silica silt, “100S”
(SIL-CO-SIL 250 from U.S. Silica). The manufacturing process to create the silt-sized particles
by crushing larger silica particles results in highly angular particles with a median particle size of
approximately 50 microns. Additional details on this soil are provided in Price (2018).
Specimen preparation. Specimens for consolidation and direct simple shear laboratory
testing were prepared with a slurry deposition method (Krage et al. 2019). A slurry of the silt is
mixed under vacuum using mechanical rotation at a water content of 29.6%, then deposited
directly into the test mold. The deposition method aims to approximate the soil fabric of fluvial
deposition. A similar deposition method was used for the centrifuge model, with mixing and
deposition scaled up by using a modified cement mixer under vacuum. 1D compression tests
were performed on air-pluviated specimens. Additional specimen preparation details are
available in Price (2018).
1D compression and consolidation tests. The results of constant rate of strain consolidation
loading for eighteen 100S specimens are shown in Figure 1a (note that several of the test
specimens were unloaded to apply a target overconsolidation). These data were collected during
the initial consolidation stage for direct simple shear tests. The test results show a range of initial
measured void ratio (eo) at 100 kPa vertical effective consolidation stress of about 0.60 to 0.63.
1D compression tests characterize compression behavior at low and high-stress conditions.
Soil compression behavior in the MIT-S1 constitutive model is based on the Pestana and Whittle
(1995) compression model and limiting compression curve (LCC). When soils are loaded to
large enough compressive stresses, the behavior falls onto the LCC regardless of initial state. For
clay soils, the LCC is analogous to the virgin compression curve. For granular soils, the LCC has
a maximum void ratio and becomes parallel to the CSL at high stresses. The laboratory-

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 375

characterized LCC and transition onto the LCC are key parts of calibrating MIT-S1. One-
dimensional compression tests to 140,000 kPa were performed to characterize the LCC and
transition onto the LCC (Haugaard et al. 2018). Results from one test for the non-plastic silt are
shown in Figure 1b. The data indicate low initial compressibility of the silt to a vertical effective
stress of about 10,000 kPa. From 10,000 kPa the compression behavior begins to transition to the
LCC regime, appearing to fully transition close to 100,000 kPa.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Laboratory test data for non-plastic silt (100S): (a) 1D consolidation tests; (b)
high-stress 1D compression test; (c) undrained direct simple shear test.

Undrained direct simple shear tests. Equivalent undrained (constant volume) direct simple
shear (UDSS) tests were performed on slurry-deposited normally consolidated silt specimens.
These tests characterized the shear behavior of the non-plastic silt and were subsequently used to
inform the calibration of the MIT-S1 model. UDSS tests were performed for three initial vertical
effective stresses (σ’vo): 100 kPa, 825 kPa, and 1680 kPa. Testing over this range of σ’vo was
advantageous to calibrate MIT-S1 over a range of stresses that are similar to those imparted by
the penetrating cone. The UDSS shear stress-shear strain test data are shown in Figure 1c. The
test data for all σ’vo conditions show a dilative response with continual hardening over the tested
shear strain range (to 15% shear strain). This highly dilative behavior (exhibited for slurry
deposited initial conditions) is likely attributable to the silt particles being highly angular.
Centrifuge model with in-flight cone penetration. In-flight cone penetration testing was
performed on geotechnical centrifuge models. Saturated soil models were constructed with
slurry-deposited non-plastic silt, consolidated under static weight, then tested at 80g on the 1 m
radius centrifuge at the University of California, Davis Center for Geotechnical Modeling. The
penetrometer advancement rate was varied between profiles to capture a range of penetration
drainage conditions. The qt data from the penetration profiles at σ’vo = 80 and 100 for average
penetration velocities from 10-5 m/s to 0.15 m/s are summarized in Figure 2a. The qt values are
corrected for porewater pressure acting on the cone shoulder and temperature effects (Price et al.,
2019a). Figure 2b shows the same cone penetration testing data as Figure 2a with qt normalized
for overburden stress (Q) versus the normalized penetration velocity (V) as defined by DeJong &
Randolph (2012). It is notable that the measured qt for 100S increases as penetration velocity
increases resulting in a qt that is larger for undrained than drained conditions. For many fine-
grained, near-normally consolidated soils, the opposite is observed (Silva 2005, Schneider et al.
2007, Kim et al. 2008). This study examines the shear dilatancy behavior of 100S as an
explanation for the observed drained and undrained qt values.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 376

A characteristic curve that describes the relationship between Q and V (DeJong & Randolph
2012) is fit to the data in Figure 2b. The penetration data and characteristic curve indicate that
drained to nearly-undrained responses were achieved for 100S by varying penetration velocity. A
drained response is achieved for penetration velocities less than about 0.01 m/s or V less than 3.
Undrained penetration conditions appear to occur for penetration velocities greater than 0.15 m/s
or V greater than about 45. The characteristic curve fit indicates that the fully undrained Q was
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

likely slightly larger than the measured Q at V = 45.


Further testing of non-plastic silt and kaolin clay mixtures in Price et al. (2017) shows that
the addition of a small amount of clay notably changes the soil behavior and response. Figure 2
includes the results of in-flight penetration for a mixture of 80% non-plastic silt and 20% kaolin
clay (80S20K) with a plasticity index of 6. The addition of kaolin clay corresponds to a notable
decrease in penetration resistance (by over an order of magnitude) for both drained and
undrained conditions compared to 100S. Additionally, the 80S20K mixture exhibits a more
typical normally consolidated fine-grained soil response to penetration drainage conditions,
where qt decreases as drainage conditions transition from drained to undrained conditions.

Figure 2. In-flight cone penetration test results from geotechnical centrifuge models.

CONE PENETRATION MODEL

Numerical cone penetration simulations were performed for both 100S and 80S20K soil
mixtures to examine the role of soil dilatancy and initial state on qt. The objective of this study is
to investigate the penetration behavior of non-plastic silt, therefore, the focus of the constitutive
model calibrations and numerical simulations is the 100S soil mixture. Further information on
the characterization and constitutive model calibration for the 80S20K mixture is available in
Price (2018).
Axisymmetric direct penetration model. Cone penetration is simulated in the finite
difference software FLAC 8.0 (Itasca 2016) with an axisymmetric direct penetration model.
Cone penetration is simulated for a single point in the soil profile, from a wished-in-place initial

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 377

condition, until steady-state cone penetration resistance, stress conditions, and porewater
pressure conditions are achieved. Typically, 10 cone diameters of simulated penetration are
required for steady-state penetration resistance. Large deformations around the penetrating cone
are accommodated with a user-defined rezoning and remapping algorithm based on an arbitrary
Lagrangian Eulerian (ALE) approach (Moug 2017, Moug et al. 2019a).
The direct penetration model geometry and boundary conditions are shown in Figure 3. The
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

boundary at x = 0 has axisymmetric boundary conditions that does not allow displacement in the
x-direction. In situ total vertical stress conditions are applied at the bottom and far radial
boundaries. An infinite elastic boundary is in place at the far radial boundary to respond to small
strains. Penetration is simulated as soil flowing up relative to a static cone by applying the
penetration velocity to the top boundary. The cone geometry is defined by interface elements.
The interface elements also capture the roughness between the cone penetrometer and soil. The
roughness coefficient used for 100S, defined as the ratio of the interface friction angle to the
critical state friction angle, is 0.20.

Figure 3. Direct axisymmetric cone penetration model in FLAC.

MIT-S1 constitutive model. The MIT-S1 constitutive model is a bounding surface plasticity
model developed by Pestana and Whittle (1999). The model is capable of capturing soil behavior
ranging from sedimentary clays to clean sands (Pestana and Whittle 2002; Pestana et al. 2002).
MIT-S1 was implemented as a user-defined constitutive model for FLAC by Jaeger (2012) and
later integrated with the user-defined ALE rezoning and remapping algorithm Moug (2017). The
MIT-S1 model was used to simulate undrained direct cone penetration in clay in Moug et al.
(2019a) and drained sand in Moug et al. (2019b). These two studies demonstrated the capability
of the penetration model with MIT-S1 to study cone penetration for these two soil types. The
present study builds from these studies by examining cone penetration in an intermediate soil.

CALIBRATION OF MIT-S1 FOR CONE PENETRATION IN 100S

Model parameter selection. Calibration of the MIT-S1 constitutive model generally


followed the approach in Moug et al. (2019b). This approach prioritizes the soil properties and
behaviors that most strongly affect qt, including the position of the critical state line (CSL) at

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 378

high stress conditions, shear strength and stiffness, and state-dilatancy behaviors. The approach
recognizes laboratory test data may not be fully captured by the calibrated model. Instead,
calibrations prioritized reasonably representing the range of soil behaviors most affecting cone
penetration, relying in some cases on well-established empirical correlations, with the aim of
capturing similar cone penetration behavior as observed in validation data.
Three calibrations for 100S were tested in this study. The MIT-S1 model parameters for each
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

calibration are summarized in Table 1. 14 of the 16 MIT-S1 model parameters were consistent
between the three calibrations. The two model parameters that differed between calibrations
primarily controlled the state-dilatancy behavior and position of the CSL. Additional details on
the approach and simulated behaviors of Calibration 1 are provided in Price (2018).

Table 1. MIT-S1 model parameters for 100S calibrations

MIT-S1 Calibration
Parameter Description
Parameter 1 2 3
𝜌𝑐 Slope of limiting compression curve in log⁡(𝑒) − log⁡(𝑝′ ) 0.33

𝜎𝑣,𝑟𝑒𝑓 /𝑝𝑎𝑡𝑚 Reference 𝑝′ at 𝑒 = 1 on the 1-D LCC 60.0
𝜃 Controls transition to limiting compression curve 0.60
𝐷 Characterizes slope of unloading curve 0.0
𝑟 Characterizes shape of unloading curve 0.0
Lateral earth pressure coefficient at normally consolidated
𝐾0𝑁𝐶 conditions 0.50
𝜇𝑜′ Small strain Poisson’s ratio 0.23
𝜔 Controls non-linearity in Poisson’s ratio 1.0
𝐶𝑏 Controls small strain elastic moduli. 899.0

𝜙𝑐𝑠 Critical state friction angle 33.0

𝜙𝑚𝑟 Peak friction angle at 𝑒 = 1 11.50 2.246 0.556
𝑝𝜙 Controls variation of peak friction angle with void ratio 3.30 7.0 10.0
𝑚 Controls shape of yield and bounding surfaces 0.45
𝜔𝑠 Controls non-linearity of elastic moduli in shear 4.0
𝜓 Controls rate of evolution of the yield surface anisotropy 60.0
Controls plastic strain magnitude when over consolidation
ℎ ratio > 1 2.0

MIT-S1 compression behavior is controlled by Cb, ρc, and σ’v,ref/patm. These parameters are
consistent between the three calibrations and were selected to capture the 1D compression
laboratory data (Figure 1b). A comparison of the simulated MIT-S1 1D compression behavior
for all three calibrations to the laboratory data is shown in Figure 4a. MIT-S1 was not able to
capture the full compression behavior of 100S observed in laboratory testing. Therefore, the
calibration prioritized the compression behavior at higher stresses.
Calibration of MIT-S1 for shear behavior aimed to approximate the trends observed in UDSS
laboratory test data. There were two areas of focus for the calibrations: (1) normally consolidated
stress-strain behavior, which was prioritized in Calibration 1, and (2) state-dilatancy behavior
across a range of initial states, which was prioritized in Calibrations 2 and 3. The results of single
element UDSS simulations are compared to the laboratory test results in Figure 4b. The initial
conditions for the single element simulations had the same σ’vc as the laboratory tests. Initial
states (ξo) were consistent with the estimated ξo in the laboratory tests, where ξo is defined as eo –
ecsl. Initial states for the laboratory tests were estimated from the position of the CSL for

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 379

Calibration 1 and were -0.02, -0.04, and -0.04 for σ’vc of 100, 825, and 1680 kPa, respectively.
Calibration 1 provides the strongest agreement with the laboratory data, while Calibrations 2 and
3 show a more strongly dilative response.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Comparison of laboratory test data and MIT-S1 single element simulations:
(a) 1D compression; (b) undrained DSS.

Figure 5a,b presents simulated and experimental state-dilatancy behaviors for the 100S silt
together with empirical relationships in the literature for non-plastic silt and clean sand. Figure
5a shows the difference between peak and critical state friction angles (ϕ’peak – ϕ’cs) vs. ξo and
Figure 5b shows ϕ’peak – ϕ’cs vs. peak dilation angle (ψmax). The experimental data for the 100S
and 80S20K mixtures (with error bars representing uncertainty) are from drained DSS tests
where denser ξo were achieved by drained cyclic loading or repeated cycles of undrained cyclic
loading and reconsolidation (Price 2018). The laboratory and simulation data are compared to
published relationships for sand (Bolton 1986, Been & Jeffries 1985, Vaid & Sasitharan 1992)
and non-plastic silt (Penman 1953) for triaxial compression (TC) and plane strain (PS)
conditions. Laboratory tests for the 100S and 80S20K mixtures exhibit greater values of ϕ’peak
– ϕ’cs for a given ξo than the published relationships and similar values of ϕ’peak – ϕ’cs for a
given ψmax as Bolton’s relationship for PS, indicating that these silts are more dilative than
clean sands.
The single element simulations shown on Figure 5a,b for the non-plastic silt are for TC,
although simulated state-dilatancy behaviors exhibited very little Lode angle dependence (e.g.,
differences between TC and PS conditions). The state-dilatancy behavior of Calibration 1 was
calibrated to be consistent with published relationships for clean sand (roughly in the middle of
Bolton’s relationships for TC and PS). Calibrations 2 and 3 aimed to be more consistent with the
laboratory test results than Calibration 1, although these calibrations still underpredict the lab
data for very small negative states (e.g., ξo ≥ −0.025). Calibration 3 has the most strongly dilative
behavior.

COMPARISON OF SIMULATED qt WITH CENTRIFUGE MEASUREMENTS

Drained and undrained penetration was simulated for the 100S calibrations. The initial
conditions were σ’vc = 100 kPa and K0 = 0.5. Penetration was simulated for ξo = -0.02 and -0.09.
The best estimate of ξo for the centrifuge model is -0.02 based on estimates of eo from the lab

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 380

element tests and ecsl from Calibration 1. Simulations for ξo = -0.09 were also performed to
capture the uncertainty in ξo and span a reasonable range of ϕ’peak – ϕ’cs (Figure 5a). Uncertainty
in ξo is associated with (1) the range of measured eo for slurry deposited lab element tests (Figure
1a), (2) scaling of the slurry deposition method from lab element tests to the centrifuge model,
and (3) uncertainty in position of the critical state line at low stresses. Additionally, at small
negative states (e.g., -0.02) the MIT-S1 calibrations underpredict the experimental state-
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

dilatancy data (Figure 5a). The dilatancy (and corresponding ϕ’peak – ϕ’cs values) of the 100S
in the cone penetration simulations can be increased by modifying the calibration (e.g.,
Calibration 1 vs. 2) or by decreasing ξo. For these reasons the authors believe that the range of
ξo used for the simulations (-0.02 to -0.09), together with the three alternative calibrations, is
reasonable.

Figure 5. State-dilatancy relationships for MIT-S1 calibrations for 100S.

Figure 6a compares undrained qt from the centrifuge models and simulations. Simulated
undrained qt at ξo = -0.02 for all three calibrations notably underestimate the centrifuge data.
Simulated results for ξo = -0.09 with Calibration 3 show the best agreement with the centrifuge
data, followed by Calibration 2, while Calibration 1 notably underestimates the centrifuge data.
Figure 6b compares drained qt from the centrifuge models and simulations. Simulated drained
qt for ξo = -0.02 underestimates the centrifuge data for all three calibrations. At ξo = -0.09, all
three calibrations yield similar drained qt values that are comparable to the measured
centrifuge data.
Overall, the simulated results from Calibration 3 most closely capture the qt data observed in
the centrifuge models. This appears to be strongly related to the state-dilatancy behavior, which
was prioritized for Calibrations 2 and 3. Additionally, Calibrations 2 and 3 were able to capture
the increase in qt as drainage transitions from drained to undrained, indicating that this behavior
is related to soil dilativity.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 381
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 6. Comparison of simulated and experimental cone penetration tip resistance for (a)
undrained and (b) drained penetration conditions.

DISCUSSION

Differences between the calibrations are attributed to state-dilatancy behavior and position of
the CSL. The two MIT-S1 parameters used to calibrate the state-dilatancy behavior (pϕ and ϕ’mr)
also affect the position of the CSL. The CSLs for the three calibrations in TC loading are shown
in Figure 7a. The calibrations were developed to have similar eCSL at low stresses (i.e., mean
effective stress, p’ = 100 kPa), then the position of the CSLs diverge at larger stresses.
Calibration 3’s CSL is positioned at the largest p’ values and Calibration 1 at the lowest.

Figure 7. Simulated e – p’ paths from in-situ conditions to just below the cone shoulder
during cone penetration compared to CSLs for MIT-S1 100S calibrations.

The position of the CSL between the three calibrations has a direct impact on the stress
conditions around the penetrating cone. Soil near the penetrating cone is loaded to conditions on
the CSL, therefore, simulated qt values are strongly influenced by CSL position. Figure 7b shows
the path of soil adjacent to the cone during drained penetration for ξo = -0.02 and -0.09. For all
calibrations and ξo, soil dilates to the CSL, reaches critical state conditions at a point ahead of the
cone tip, then continues loading along the CSL to the cone shoulder. The cone tip is indicated in
Figure 7b for ξo = -0.09 simulations. The final p’ conditions at the cone tip relate to the simulated

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 382

qt values. Given the position of the CSLs, a stronger dilation response is observed from
Calibrations 2 and 3 to reach the CSL during drained penetration. In the case of undrained
penetration, shown in Figure 7c, there is no change in e to reach the CSLs. Therefore, the
differences in simulated qt largely reflect the relative p’ of the CSLs for the same eo.
The simulated qt for ξo = -0.09 with Calibrations 2 and 3 is larger for undrained conditions
than drained conditions; this is consistent with the centrifuge-measured qt values. This response
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

is attributed to a strongly dilative response and relates to both the CSL position and ξo. For both
Calibrations 2 and 3 at ξo = -0.09, the simulated drained responses have strong dilation to the
CSLs at low p’, then continue to larger p’ values at the cone face along the CSL. For undrained
penetration, the increase in p’ to the CSL is very large due to the CSL being positioned at high
p’, resulting in larger p’ values at the cone face than for the drained response.
The measured and simulated qt for 100S are compared to other soils for context and further
insight into how compressibility, state, and dilatancy affect soil behavior during cone
penetration. The measured and simulated penetration behavior of the 80S20K mixture are
compared to the results of 100S in Figure 6a,b. The 80S20K centrifuge testing and MIT-S1
calibration are described in Price et al. (2019) and Price (2018), respectively. The 8020SK
mixture exhibits undrained shear strength normalization and contractive tendencies during
undrained shearing for slurry deposited states and is much more compressible than the 100S
(Price et al. 2019). Note that the data points for the 80S20K mixture on Figure 5a,b (which
follow a similar trend as the 100S data) are for ξo much denser than can be achieved for normally
consolidated slurry deposited specimens. The 8020K simulations for ξo of 0.05 (K0 normally
consolidated, slurry deposited condition) show close agreement to the centrifuge-measured qt for
drained and undrained penetration. The 80S20K simulations for ξo = -0.02 and -0.09 produced
undrained qt approximately 2 and 5 times larger, respectively than for ξo = 0.05; and drained qt
approximately 1.5 and 1.9 times larger, respectively. Calibration 1 for the 100S and the 80S20K
calibration used herein exhibit similar state-dilatancy behaviors for small negative states (Price
2018). Therefore, differences between simulated qt for these materials with ξo = -0.02 largely
reflect differences in compressibility, which affects position of the respective CSLs as described
in Price (2018).
In addition, the relationship between simulated drained qt and ξo for Ottawa sand from Moug
et al. (2019b) is shown in Figure 6b. Overall, the measured qt for the 100S (for slurry deposited
conditions) are greater than for drained Ottawa sand (relative densities up to 60% are shown on
Figure 6b, however, the slurry deposited drained qt for the 100S is greater than Ottawa Sand for
even a relative density of 80% based on data from Moug et al. 2019b) and more than an order of
magnitude greater than for the 80S20K mixture (both drained and undrained conditions). Within
this context, the simulations for Calibration 2 and 3 and ξo = -0.09 reasonably capture the
experimental data for the 100S and demonstrate the ability of the numerical cone penetration
model to represent a wide range of soil behaviors.

CONCLUSIONS

Cone penetration in a non-plastic manufactured silt was examined with geotechnical


centrifuge modeling and direct numerical simulations of cone penetration. Simulated cone
penetration was performed for three different MIT-S1 constitutive model calibrations. The
primary difference between the constitutive model calibrations was the state-dilatancy behavior,
which was also reflected in differences in the position of the CSLs at high p’ conditions. Cone

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 383

penetration simulations with the most strongly dilative calibration, Calibration 3, captured
similar drained and undrained qt values as measured in the geotechnical centrifuge model for ξo =
-0.09.
The simulated cone penetration results demonstrated that state-dilatancy relationships
strongly affect qt for drained and undrained cone penetration. The 100S soil exhibited a very
strongly dilative response which resulted in higher drained qt than for a clean silica sand (i.e.,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Ottawa sand) for similar ξo. Constitutive model calibrations for studying cone penetration should
prioritize the soil’s state-dilatancy behavior and position of the CSL at high stresses, which can
be partly constrained by high stress 1D compression testing to measure the soil’s LCC. State-
dilatancy relationships should be incorporated into theory-based interpretation of CPT data for
intermediate soils.
The study demonstrated that cone penetration responses were reasonably well captured for
this highly dilative, manufactured non-plastic silt. This soil exhibited behaviors that are not
likely to be observed for natural non-plastic silts, including strongly dilative behaviors for small
initial negative states (achieved by slurry deposition). These results further indicate the utility of
the numerical cone penetration model with MIT-S1 for representing a wide range of soil types,
including natural non-plastic silts, and for developing a theory-based understanding of cone
penetration in intermediate soils including non-plastic silts.

ACKNOWLEDGEMENTS

The authors appreciate the support of the University of California, Davis SILab, Center for
Geotechnical Modeling, and Portland State University to perform this work.

REFERENCES

DeJong, J. T., and Randolph, M. (2012). Influence of partial consolidation during cone
penetration on estimated soil behavior type and pore pressure dissipation measurements. J.
Geotech. Geoenvir. Eng., 138(7), 777-788.
Krage, C. P., and DeJong, J. T. (2016). Influence of drainage conditions during cone penetration
on the estimation of engineering properties and liquefaction potential of silty and sandy soils.
Journal of Geotechnical and Geoenvironmental Engineering 142(11), 04016059.
Krage, C. P., Price, A. B., Lukas, W. G., DeJong, J. T., DeGroot, D. J., and Boulanger, R. W.
(2019). Slurry deposition method of low-plasticity intermediate soils for laboratory element
testing. Geotechnical Testing Journal, 43(5), 1269-1285.
Haugaard, S., Price, A., DeJong, J., and Boulanger, R. (2018). “One-dimensional compression
testing for low-plasticity silts”, in One-dimensional compression testing for low-plasticity
silts. DesignSafe-CI.
Houlsby, G. T., and Hitchman, R. (1988). Calibration chamber tests of a cone penetrometer in
sand. Geotechnique, 38(1), 39-44.
Itasca. (2016). FLAC–Fast Lagrangian Analysis of Continua, Version 8.0. Minneapolis: Itasca
Consulting Group.
Jaeger, R. A. (2012). Numerical and experimental study on cone penetration in sands and
intermediate soils. Ph.D. thesis, Dept. of Civil & Environmental Engineering, Univ. of
California, Davis.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 384

Moug, D. M. (2017). Axisymmetric cone penetration model for sands and clays. PhD
Dissertation, Dept. of Civil & Environmental Engineering, Univ. of California, Davis.
Moug, D. M., Boulanger, R. W., DeJong, J. T., and Jaeger, R. A. (2019). Axisymmetric
simulations of cone penetration in saturated clay. Journal of Geotechnical and
Geoenvironmental Engineering, 145(4).
Moug, D. M., Price, A. B., Parra Bastidas, A. M., Darby, K. M., Boulanger, R. W., and DeJong,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

J. T. (2019). Mechanistic development of CPT-based cyclic strength correlations for clean


sand. Journal of Geotechnical and Geoenvironmental Engineering, 145(10).
Pestana, J. M., and Whittle, A. J. (1995). Compression model for cohesionless soils.
Géotechnique, 45(4), 611-631.
Pestana, J. M., and Whittle, A. J. (1999). Formulation of a unified constitutive model for clays
and sands. International Journal for Numerical and Analytical Methods in Geomechanics,
23(12), 1215-1243.
Pestana, J. M., Whittle, A. J., and Salvati, L. A. (2002a). Evaluation of a constitutive model for
clays and sands: Part I–sand behaviour. International journal for numerical and analytical
methods in geomechanics, 26(11), 1097-1121.
Pestana, J. M., Whittle, A. J., and Gens, A. (2002b). Evaluation of a constitutive model for clays
and sands: Part II–clay behaviour. International journal for numerical and analytical
methods in geomechanics, 26(11), 1123-1146.
Price, A. B., DeJong, J. T., and Boulanger, R. W. (2017). Cyclic loading response of silt with
multiple loading events. Journal of Geotechnical and Geoenvironmental Engineering,
143(10), 04017080-04017080.
Price, A. B. (2018). Cyclic strength and cone penetration resistance for mixtures of silica silt
and kaolin. PhD dissertation, Department of Civil & Environmental Engineering, University
of California, Davis.
Price, A. B., Boulanger, R. W., and DeJong, J. T. (2019). Centrifuge modeling of variable-rate
cone penetration in low-plasticity silts. Journal of Geotechnical and Geoenvironmental
Engineering, 145(11), 04019098.
Schnaid, F., Bedin, J., and Costa Filho, L. M. (2010). Drainage characterization of tailings from
in situ test. In Proceedings of the 2nd International Symposium on Cone Penetration Testing,
Robertson, Huntington Beach, CA, USA.
Schneider, J. A., Lehane, B. M., and Schnaid, F. (2007). Velocity effects on piezocone tests in
normally and overconsolidated clays. Int. J. Phys., 7(2), 23-34.
Silva, M. F. (2005). Numerical and physical models of rate effects in soil penetration. PhD
dissertation, University of Cambridge.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 385

Creep, Relaxation, and Strain Rate Effects in Central Florida Silty Sand

Sergio Marin1 and Luis G. Arboleda-Monsalve2


1
Research Assistant, Dept. of Civil, Environmental, and Construction Engineering, Univ. of
Central Florida, Orlando, FL. Email: marinsergio81@knights.ucf.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Assistant Professor, Dept. of Civil, Environmental, and Construction Engineering, Univ. of
Central Florida, Orlando, FL. Email: Luis.Arboleda@ucf.edu

ABSTRACT

Time and rate dependency studies of soil materials have been focused on clayey soils where
the viscous effects are more pronounced for low confining pressures. In sandy materials, time-
dependent and rate effects are usually considered negligible. Therefore, for practical purposes
sands are deemed as non-viscous materials although there is experimental evidence that creep
strains of 10% under monotonic loading can be developed. The viscous behavior of granular
materials has been demonstrated to be stress dependent. For low confining pressures, it is attributed
to particle rearrangement. Most studies on compressibility time-dependent effects of granular soils
were developed in one-dimensional or triaxial conditions. The effects of different contents of non-
plastic silts have been studied showing that granular soils with fine contents around the transitional
limit of approximately 35%–45% start to behave in compressibility testing more like a fine grained
material following isotach characteristic behaviors. The objective of the testing program presented
herein, which was performed in a constant rate of strain apparatus for low confining pressures, was
to evaluate the effects of fines in the material compressibility response of Central Florida silty
sands. Control tests on several relative densities were targeted for clean sands. Creep, relaxation,
and strain rate effects were evaluated in soil samples reconstituted to several effective axial stress
conditions, and fines were added to the clean sand to target void ratios corresponding to the loose
state. In this research, the concept of global void ratios versus skeleton void ratios was used to
compare the silty sand behaviors up to a 20% of fine content in relation to control tests conducted
in clean sands. This research concluded on the relative importance of considering the participation
of fines up to 20% weight based on the compressibility response for several sample preparation
techniques used to reconstitute the soil samples in the laboratory.

INTRODUCTION

Leroueil (1996) stated that viscous effects in soils are not negligible although in practice they
are ignored particularly in granular materials. Strain rates and temperature effects on soil behavior
were also described by Leroueil (1996) and it was found that in clays there are changes of up to
10% in undrained shear strength and preconsolidation pressure for each logarithm cycle of strain
rate, which is 2 to 4 orders of magnitude smaller in the field than in the laboratory. Tatsuoka et al.
(2008) presented four viscosity types of behavior: isotach, combined, temporary or transient
effects of strain rate, and strain acceleration (TESRA), and positive and negative (P & N). The
most common viscosity type is the isotach which is characterized by an increase in the viscous
stress (V) due to a step increase in the strain rate (𝜀̇). The other three types experience a decay in
V during the monotonic loading. TESRA behavior shows an increment in V to return to the
initial compressibility curve. Combined behavior shows a smaller decay than the initial V, and

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 386

for the P & N behavior the decay is larger than the initial V shifting the curve towards a smaller
viscous stress. The same authors separated the time effects of the geomaterial in ageing (i.e.,
changes over time in the stress strain properties such as yield, peak strength, among others), and
loading rate effects (i.e., stress strain behavior changes due to strain rate on monotonic load without
including the effects of delayed dissipation of excess pore water pressure). In addition, Augustesen
et al. (2004) defined creep as a process characterized by an increase in strain at a constant stress
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

which under triaxial conditions was further divided into: 1) primary creep or transient creep, 2)
secondary creep or stationary creep, and 3) tertiary creep or acceleration creep. Each creep stage
is differentiated in a log-time versus log-strain rate plot where the strain rate initially decays
(primary), it stays constant (secondary), and finally it increases (tertiary). For oedometric tests,
Augustesen et al. (2004) concluded that only the primary creep can be observed. Also, relaxation
was described by Ladanyi et al. (1995) as the stress reduction while a given level of strain is held
constant. Ladanyi and Benyamina (1995) stated that a basic difference between creep and
relaxation is the material structure changes with time which for relaxation remains fixed leading
to a redistribution of the internal stresses. In addition, Augustesen et al. (2004) stated that granular
materials reveal larger rheologic deformations when there is an exposure to high confining
pressure which is caused by particle crushing more than particle rearrangement. Lade et al. (1996)
described that particle breakage is a function of time, which depends on the material mineralogy,
and that creep is its external manifestation. Lade et al. (2016) performed sieve analysis on samples
before and after creep and relaxation. The results showed that particle crushing during relaxation
is less pronounced than for creep. Levin (2021) studied the time-dependent effects of four different
sands and the influence of non-plastic silts content up to 25%. It was found that for fine contents
larger than 14%, creep can be described with the viscosity index (Iv), which is the ratio between
creep and compression indexes, C and Cc respectively. Herein, the sand and silty sand materials
were tested under one-dimensional compression using a constant rate of strain apparatus with
similar variables and conditions as by Levin (2021). A silty sand with a FC of 20%, which is a
typical percentage in Central Florida, was considered in the testing program. Thus, the strain
history independence, rate dependance, relaxation, and creep effects were linked to Iv to evaluate
the effects of fines on the compressibility response.

MATERIAL CHARACTERIZATION

The material used in this experimental program was retrieved from two boreholes, labeled as
B1 and B2, using an auger extended to depths ranging from 1.5 to 6 meters under the ground
surface at a testing site near the University of Central Florida main campus. A natural moisture of
approximately 20% was determined after oven drying three randomly selected portions for each
depth. The oven dried material was washed to determine the natural fine content (FC), which was
approximately 15%, and it was kept clean to determine the gradation curve as presented in Figure
1. Specific gravities tests were performed on the coarse material (GC) and fine portion (Gf) with
values of 2.65 and 2.48, respectively. Each material was tested independently to compare their
compressibility before assessing their shear behavior in a triaxial environment which is part of the
experimental program to study the effects of fines for different fine contents in Florida Sands.
The packing characteristics of the tested soil, represented by the maximum and minimum void
ratios (i.e., emax and emin), were determined in the laboratory using for emax the procedure proposed
by Lade et al. (1998). emin was determined using a proctor mold and compacting the clean sand in
five layers with 25 blows of a Marshall hammer. The nominal values were: emax=1.03 and
emin=0.58. Published correlations with D50 and particle shape were used for verification. Images

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 387

collected with a light microscope were used to determine the particle shape (i.e., roundness and
sphericity) of the material retained on the sieve #80 (more than 50%). From the randomly selected
particles and using an average value, the material was classified with a sphericity (S) of 0.58, a
roundness (R) of 0.42, and a Regularity (Reg.) of 0.50. Cho et al. (2006) compiled material
properties from published studies to find correlations between particle shape and soil properties
including information of the extreme void ratios which were also used for verification. These two
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

extreme values allow to approximate a relative density (Dr) of the testing material which is only a
reference value more than a state parameter. The effect of the FC on the extreme void ratios was
studied by Lade et al. (1998), Cubrinovski et al. (2002), among others.

Figure 1. Grain size distribution of the clean sand

The effects of the FC on the soil behavior has been studied by Yang et al. (2006) and Zuo et
al. (2015) which provided insight into the transitional FC that separates the behavior between sand-
dominated and fine-dominated. Eq. 1 was presented by Zuo and Baudet (2015) to determine the
skeleton void ratio (es) in studies where the behavior was sand-dominated. Eq. 1 is further
simplified when fine and coarse materials have the same specific gravities.

𝑉𝑉 + 𝑉𝑓 𝑒(𝐺𝑓 − 𝐺𝑓 𝐹𝐶 + 𝐺𝐶 𝐹𝐶) + 𝐺𝐶 𝐹𝐶
𝑒𝑠 = = (1)
𝑉𝑆 𝐺𝑓 (1 − 𝐹𝐶)

The global void ratio, e in Eq. 1, considers the fine particles that are not part of the structural
skeleton or are moved to the voids at small stresses. When the behavior is fine-dominated, the sand
particles are ignored in the force chain (Zuo and Baudet 2015) and Eq. 2 is used to calculate the
interfines void ratio (ef)

𝑉𝑉 𝑒(𝐺𝑓 − 𝐺𝑓 𝐹𝐶 + 𝐺𝐶 𝐹𝐶)
𝑒𝑓 = = (2)
𝑉𝑓 𝐺𝐶 𝐹𝐶

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 388

This concept is required to determine the limiting fine content, also called transitional fine
content (TFC), where the fines start to participate in the boundaries around the coarse particles
instead of just filling their void space.

SPECIMEN PREPARATION AND TESTING PROCEDURES


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Dry pluviation (DP) and moist tamping (MT) methods, as described by Murthy et al. (2007),
were used for the sample preparation. The difference with Murthy et al. (2007) recommendations
was that instead of using 8% of water content and 6 to 10 layers for sample reconstitution in triaxial
specimens, a range of water content between 5 and 8% was used for the clean sands, which
increased up to 12% for the 20% of FC specimens. The number of layers ranged from 4 to 8. A
funnel was used to deposit the material for the DP. After all the material was placed in the CRS
oedometer ring, the soil was gently flattened from its conical shape with a metallic ruler using the
top edge of the ring as reference. Moist pluviation (MP) was also used following the same
procedure but changing the moisture content to less than 1%. That small portion of water provided
stability to the material during sample preparation allowing to reach higher void ratios close to
emax. The 20% FC mixture did not allow to reach low skeleton void ratios. The ratio of FC was
weight-based for all the samples tested. The initial void ratios achieved during preparation were
initially higher for the loose samples. The saturation stage reduced the void ratios towards medium
dense materials. The loosest stage was attempted to be achieved for later comparison with the
triaxial specimens which are part of future work on this experimental program. The medium and
dense samples experienced less changes during the saturation stages. Medium dense samples are
the ones mostly found in Central Florida. Thus, the samples corresponding to such classification
are characteristic for this location. Dense samples are also a reference for comparison with the
future work on the triaxial tests.
All the specimens were tested under strain-controlled conditions under constant rate (CRS).
The strain rates selected were 1, 5, 6, 10, and 40%/hr. Initially, a single strain rate was used for
each sample until an effective axial stress of approximately 2000 kPa was reached. All the
specimens were subjected to creep at the end of the loading stage before unloading the specimen
to consider aging and to avoid any inertial effect. Additional creep stages were performed at
approximately 200, 500, and 1000 kPa. For all the creep stages, a constant stress module in the
GDS software was used. That module allows the user to keep constant the effective or total axial
stress and the back pressure. The creep stages were initially held for at least 90 mins. The creep
time was reduced to times where the Creep Axial Loading Rate (ALR) was close to the LVDT
accuracy. The relaxation stages were set up with the CRS module where the input strain rate was
zero and the back pressure was kept constant for about two hours. The stresses for relaxation were
around 250, 600, 1200 kPa, and at the end of the reloading stages for some samples. Combinations
of creep and relaxation on the same samples were also performed such that after the creep stage,
the initial strain rate was used until the initial isotache curve was reached again. Table 1 presents
a summary of the sample preparation methods, FC, strain rates, initial void ratios (global and
skeleton) and relative densities, stresses at creep, and initial stresses for relaxation.

EXPERIMENTAL RESULTS

The results of strain rate dependency, creep and relaxation are presented herein. The void ratio
presented for each sample was tracked during docking, saturation, and pseudo B-value check

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 389

stages in the CRS device. Regardless of the deposition method, for high void ratios initial load
setting stages were observed in the very small axial strains for the clean sand, which was corrected
both in terms of loading and void ratios, only when that occurred at loading ranges near the docking
load. For each sample, the nomenclature consists of the deposition method used (DP, MP, or MT),
the FC in parenthesis, the strain rate, and the void ratio including both the global (e0) and skeleton
(es) void ratios. Recall that the global and skeleton void ratios are only used for samples with fines.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Once the creep and relaxation tests were finished, the material that was reused for them was sieved
to identify particle breakage. The gradation curves for both materials (B1 and B2) remained the
same showing that the particles were not broken for the applied pressures (i.e., approximately a
maximum of 2000 kPa).

Table 1. Specimen preparation and testing procedures

Specimen FC Packing Characteristics CRS Stresses at Creep Initial Stresses


at relaxation
# Prep. % e0 eS Dr %/hr kPa kPa
(%)
S1 MT 20 0.45 0.84 42 40 2000 ----
S2 MT 20 0.57 0.99 10 5 200, 500,1000, 2000 250, 600, 1200
S3 DP 0 0.88 0.88 33 40 2000 ----
S4 DP 0 0.93 0.93 22 5 2140 ----
S5 DP 0 0.65 0.65 84 40 2000 ----
S6 DP 0 0.63 0.63 89 1 2140 ----
S7 DP 0 0.63 0.63 89 5 2140 ----
S8 MT 0 0.61 0.61 92 6 2000 ----
S9 MT 20 0.71 1.18 ---- 5 200, 500, 1000, 2000 250, 600, 1200
S10 MT 20 0.52 0.92 24 5 200, 500, 1000, 2000 250, 600, 1200
S11 MT 0 0.91 0.91 27 40 200, 500, 1000, 2000 250, 600, 1200
S12 MP 0 0.89 0.89 32 5 200, 500, 1000, 2000 225, 525, 1050
S13 DP 0 0.84 0.84 43 40 200, 500, 1000, 2000 ----
S14 DP 0 0.66 0.66 82 5 200, 500, 1000, 2000 ----

STRAIN RATE OF TESTED MATERIALS

The compressibility dependency on the strain rate was evaluated attempting to avoid changes
in the strain rate along the entire test. Nonetheless, the Isotach behavior permitted to use any other
test as long as the initial strain rate was applied after any other change in strain rate or stage (Creep
and relaxation). Figure 2 presents the compressibility response for samples with high void ratios
(a) and more denser materials (b) at different strain rates. Figure 2.a also includes materials with
20% of fine contents. In that figure, it is noticed that the shape of the silty sand has a more gradual
change in slope than the sandy material. The clean sand compressed at a higher strain rate tends to
follow that pattern at pressures below 15 kPa. After around that pressure, the strain rate on the silty
sand started to show deviations making the material with higher strain rate more compressible.
The tested clean sand also showed more compressibility for higher strain rates. Although the shape
is different, for small strain rates the axial strain tends to be the same at the final axial effective
stress (a’); the clean sand did not show a change in slope. Comparing only the silty sand samples,
the slope at the end of the test shows that for higher a’, the sample with a lower strain rate tends
to experience more axial deformation. The unloading stage for all of them show a similar slope.
Figure 2.b shows the compressibility for samples with lower void ratios, and only clean sandy soils

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 390

are presented. That figure shows that for rates up of 1, 5, and 6%/hr there is no difference in the
measured response as opposed to the case of higher void ratios that the material showed less
compressibility. The slope at the end of the loading stage shows that S5 (i.e., higher tested strain
rate) remained less compressible than the samples under smaller strain rates. Same slope is noticed
for the unloading stage.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

CREEP AFTER CONSTANT RATE OF STRAIN

The viscous behavior is represented herein by the ratio between C and Cc which is also called
viscosity index (Iv). High fine content soils should behave fine-dominated with a constant value
of Iv. To conduct these tests, the samples had to enter creep stages at different effective axial
pressures. Figure 3 presents the evaluation of Iv for silty sands at different void ratios and strain
rates. S9 shows the fine-dominated behavior with a constant ratio of 0.009. The skeleton void ratio
that is higher than emax explains that the coarse particles were not in contact and surrounded by the
fine portions. S10 does not have a constant ratio and behaved more coarse-dominated.

Figure 2. Compressibility at different strain rates and initial void ratios.

Figure 4.a and b show the evaluation of Iv for clean sands. The variation with the stress is more
pronounced for higher void ratios and higher strain rates before the material enters the creep stage.
There is not a clear convergency to a specific number for all the tested materials at higher a’.
Figure 5 presents the variation of creep and compression indexes at different stresses. C
presented in Figure 5.a shows less variation with stress. More than the strain rate, this index shows
more dependence on the packing state. The lower the void ratio, the lower the value of C and Cc.
The compressibility index (presented in Figure 5.b), shows an increase with pressure. The packing
state makes the variation of this index lower. Cc also shows dependency on the strain rate before
entering the creep stage.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 391
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. C/Cc evaluation for silty sands

Figure 4. C/Cc evaluation for silty clean sands

Levin (2021) indicated that when C/Cc is used as a descriptive parameter for the time-
dependent behavior, for lower stresses and larger densities, the material should show a larger
response. Thus, larger FC are required to guarantee that the coarse particles are surrounded by
fines.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 392
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5. C and Cc at different stress levels

Figure 6 shows the sample responses during creep at an approximate effective axial stress of
200 kPa. The initial void ratio and strain rate dependency of the creep axial loading rate (ALR) is
shown in that figure. For lower void ratios and higher rate of loading, approximately ten minutes
were required to reach the LVDT accuracy of 0.004%. All the samples required less than 30 min
to reach the LVDT accuracy. Also, the figure shows that the initial ALR is affected by the loading
strain before entering the creep stage.

Figure 6. Creep response for different samples at ’a of 200 kPa

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 393

RELAXATION AFTER CONSTANT RATE OF STRAIN

Several packing conditions and strain rates are presented. Levin (2021) indicated that small
strain rates (i.e., 0.6%/min) are recommended to reduce the load frame reaction. Nonetheless, at
higher stresses some initial strain changes were reported as a reaction of the material itself (Levin
et al. 2019). In this research, slight strain changes (i.e., 0.005%) were also recorded which are more
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

attributed to the LVDT accuracy than the reaction frame.


Relaxation in fine grained materials is evaluated with Iv, which represents the change in stress
upon a relaxation time given in Eq.3 (Levin 2021).

𝜎0′
𝑙𝑛
𝜎𝑖′
𝐼𝑉 = (3)
𝑡′
𝑙𝑛 𝑖′
𝑡0

Figure 7 shows the evaluated viscosity indexes, the compression curves in logarithmic
horizontal axis, and the stress reduction during the relaxation times. The stress reduction is not
noticeable in the compression curves, even for samples with a relaxation time larger than 24 hours.
Therefore, the quasi-isochronous behavior is not clearly represented without the aid of Figure 7.c
where the curves show some parallelism for each sample at smaller stresses than 1000 kPa. In
addition, the stress reduction is more noticeable for the silty sands than clean sands. Figure 7.b
only presents the compression curves for samples with similar strain rate and S9 is not included
because there is not a clean sand sample with such a high initial void ratio. The evaluated Iv are
shown in Figure 7.a. In that figure, a reduction of Iv is presented as the relaxation initial stress
increases. It can be noticed that the Iv reduction depends on the initial strain rate with a smaller
reduction for larger initial strain rates. Levin (2021) described that Iv for relaxation is qualitatively
similar to the Iv for creep showing that at relaxation the indexes are 1.25-1.7 times larger than
C/Cc depending on the stress level. In this research, the silty sand showed larger ratios varying
from 1.41 to 1.08. For the clean sands the relaxation Iv values were smaller than those measured
for creep (0.84-0.58) with an exception for S12 at a ’a of approximately 200 kPa where that ratio
was 1.16.

Figure 7. Relaxation analysis. a) Viscosity index versus pressure, b) compression curves,


and c) stress relaxation over time.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 394

A similar reduction of the normalized relaxation was found by Levin (2021) for a different
sand. The author presented another sand that had Iv constant. Reduction of Iv is not a general
behavioral trend for all sandy soils. The explanation provided was that relaxation stresses are
affected by the oedometric swelling index.

CONCLUSIONS
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

This paper presented the evaluation of the transitional fine content through the comparison of
the rheologic behavior of fines. The evaluation of rate dependency for the tested sands and silty
sands was presented and the main findings of the experimental program were:
• S10 with a skeleton void ratio of 0.92 and a CRS of 5%/hr showed that a FC of 20% was
not enough to reach the transitional point for the material to behave as fine-dominated.
Additional tests with larger FC are required to find the transitional point.
• Pressures applied were not large enough to reach particle breakage for the sands. It was
verified by sieving the material after all creep and relaxation tests and compare the before
and after material condition.
• Although the fine-dominated rheological behavior was not reached, the compressibility of
the material was highly affected by the FC as well as by the strain rate. When the void
ratios for clean sands are very low, the material behaved less compressible.
• Creep did not show stress dependency for the stresses up to 2000 kPa used in the
experiments. Creep was more affected by the packing state.
• The viscosity index determined by the relaxation and creep where within the same order of
magnitude being the measure for relaxation larger only for the silty sands.
• Times to achieve the maximum strain changes at creep stages were smaller compared with
the time required to reach the minimum stress at relaxation.
• The reconstitution methods used herein did not affect the compressibility curves. This
analysis was not considered for creep and relaxation stages.

ACKNOWLEDGMENTS

The Florida Department of Transportation (FDOT) financially supported this project by the
Grant No. BDV24 TWO 977-29. FDOT also provided assistance to extract the soil samples. The
opinions, findings, and conclusions expressed in this publication are those of the authors and not
necessarily those of the State of Florida Department of Transportation or the U.S. Department of
Transportation.

REFERENCES

Augustesen, A., Liingaard, M., and Lade, P. V. (2004). “Evaluation of Time-Dependent


Behavior of Soils.” International Journal of Geomechanics 4(3): 137-156.
Cho, G.-C., Dodds, J., and Santamarina, J. C. (2006). “Particle Shape Effects on Packing
Density, Stiffness, and Strength: Natural and Crushed Sands.” Journal of Geotechnical and
Geoenvironmental Engineering 132(5): 591-602.
Cubrinovski, M., and Ishihara, K. (2002). “Maximum and minimum void ratio characteristics of
sands.” Soils and foundations 42(6): 65-78.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 395

Ladanyi, B., and Benyamina, M. B. (1995). “Triaxial relaxation testing of a frozen sand.”
Canadian Geotechnical Journal 32(3): 496-511.
Lade, P. V., and Karimpour, H. (2016). “Stress drop effects in time dependent behavior of quartz
sand.” International Journal of Solids and Structures 87: 167-182.
Lade, P. V., Liggio, C., and Yamamuro, J. A. (1998). “Effects of non-plastic fines on minimum
and maximum void ratios of sand.” Geotechnical testing journal 21: 336-347.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Lade, P. V., Yamamuro, J. A., and Bopp, P. A. (1996). “Significance of Particle Crushing in
Granular Materials.” Journal of Geotechnical Engineering 122(4): 309-316.
Leroueil, S. (1996). Importance of strain rate and temperature effects in geotechnical
engineering, Measuring and Modeling Time Dependnet Soil Behaviour. ASCE, GSP 61: 1-
60.
Levin, F. (2021). “Time-Dependent Compression Behavior of Sands under Oedometric
Conditions.” Journal of Geotechnical and Geoenvironmental Engineering 147(12):
04021144.
Levin, F., Vogt, S., and Cudmani, R. (2019). “Time-dependent behaviour of sand with different
fine contents under oedometric loading.” Canadian Geotechnical Journal 56(1): 102-115.
Murthy, T., Loukidis, D., Carraro, J., Prezzi, M., and Salgado, R. (2007). “Undrained monotonic
response of clean and silty sands.” Géotechnique 57(3): 273-288.
Tatsuoka, F., Di Benedetto, H., Enomoto, T., Kawabe, S., and Kongkitkul, W. (2008). “Various
viscosity types of geomaterials in shear and their mathematical expression.” Soils and
Foundations 48(1): 41-60.
Yang, S., Lacasse, S., and Sandven, R. (2006). “Determination of the transitional fines content of
mixtures of sand and non-plastic fines.” Geotechnical Testing Journal 29(2): 102-107.
Zuo, L., and Baudet, B. A. (2015). “Determination of the transitional fines content of sand-non
plastic fines mixtures.” Soils and Foundations 55(1): 213-219.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 396

A Comparison of Approaches for Determining the Virgin Compression Line of Remolded


Saturated Soils

Alireza Shiri, S.M.ASCE1; Daniel R. VandenBerge, Ph.D., M.ASCE2; and Yousef Shiri, Ph.D.3
1
Graduate Research Assistant, Dept. of Civil and Environmental Engineering, Tennessee
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Technological Univ., Cookeville, TN. Email: ashiri42@tntech.edu


2
Associate Professor, Dept. of Civil and Environmental Engineering, Tennessee Technological
Univ., Cookeville, TN. Email: dvandenberge@tntech.edu
3
Assistant Professor, Dept. of Mining, Petroleum, and Geophysics Engineering, Shahrood Univ.
of Technology, Shahrood, Iran. Email: yousefshiri@shahroodut.ac.ir

ABSTRACT

Many advanced soil constitutive models are modifications of the Modified Cam-Clay model.
They are based on the concept of critical state theory and use several parameters to simulate the
mechanical behavior of soils. Among these, parameters describing the compression behavior of
the soil must be determined, even though one-dimensional consolidation tests may not be
available, especially for preliminary studies. This information may also be needed to switch from
a less complex soil model (e.g., Mohr-Coulomb) to a more complex one (e.g., Modified Cam-
Clay). In this paper, the slope of the virgin compression line () was estimated by advancing the
existing methods in the literature with rapid laboratory testing equipment, such as the fall cone
and miniature vane shear tests. By using this approach, two methods are proposed to estimate the
value of λ. Then, these methods were applied on three different soils, and the results were
compared and verified using one-dimensional consolidation test. The results obtained from the
proposed test methods showed an accurate estimation of the λ parameter for clays, which makes
these methods reliable, fast, and economical. But, as the clay fraction of the soil decreases, the
methods tend to overestimate the λ values.

INTRODUCTION

Modeling the mechanical behavior of fine-grained soil plays an important role in soil
mechanics (e.g., (Alamanis et al. 2021; Guo et al. 2022; Le and Airey 2021; Zhang et al. 2018)).
Critical state soil mechanics (CSSM), developed by Schofield and Wroth (1968), has long been
regarded as a suitable unified framework in which to develop geotechnical constitutive models
for modeling mechanical behavior of the soils. While initially developed for reconstituted clays,
CSSM has been widely used to explain many behavioral aspects of other different geo-materials.
Most of today’s advanced soil constitutive models are based on the CSSM-like Modified Cam-
Clay model (Roscoe and Burland 1968; Wood 1990). Determining the appropriate parameter
values remains one of the primary modeling challenges for projects with limited resources.
One of the most important CSSM parameters is λ, which is the slope of the virgin
compression line for specific volume (v) versus natural logarithm of mean effective stress (p’).
By far, the most reliable and precise method to derive this parameter is through the one-
dimensional consolidation test on natural or remolded soils. The primary disadvantage of this
method is that the consolidation test is very time consuming. Therefore, newer, faster, and
repeatable ways are being explored to determine the mechanical behavior of soils, such as the

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 397

fall cone test and miniature vane shear test (Dastider et al. 2021; Farias and Llano-Serna 2016;
Nasroulla et al. 2016). In addition to laboratory methods, correlations can be used to estimate the
virgin compression line from index tests, such as Burland’s (1990) correlation to liquid limit.
This paper compares two alternative methods for deriving λ and compares the results with the
values from the one-dimensional consolation test. To be more specific, the λ parameter will be
determined by using Burland’s intrinsic compression theory (Burland 1990) and using the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

combination of fall cone test and miniature vane shear test with the theory proposed by Wroth
and Wood (Wroth and Wood 1978). The conclusions will help to guide of the selection of
alternate means for determining the critical state parameters using fast and simple approaches.

BACKGROUND

Critical state soil mechanics was introduced by the Cambridge Soil Mechanics Group in the
1950s and 1960s (Wood 1990). Originally, the constitutive models for clays were developed
using the concept of CSSM, resulting in the original and the modified Cam-clay models (Roscoe
et al. 1963; Burland 1968). The CSSM framework assumes the existence of a critical state at
which large shearing deformation continues without dilatancy and change of the stress ratio. The
stress conditions in CSSM are described in terms of the principal stresses, using q and p’:

q =  '1 −  '3 (1)

p ' = ( '1 +  '2 +  '3 ) 3 (2)

where 𝜎1′ , 𝜎2′ , and 𝜎3′ are the effective principal stresses. In a triaxial compression test, 𝜎1′ is
vertical, and 𝜎2′ and 𝜎3′ are equal, representing the effective lateral stresses. The critical state line
(CSL) in q-p’ space describes the stress conditions associated with failure as shown in Figure
1(a) and is related to the critical state effective friction angle.

Figure 1. Critical state line in p’-q-v space (after Budhu 2020).

In addition to shear strength, consolidation properties are essential for modeling soil behavior
and are conventionally derived from one-dimensional consolidation tests. While commonly
evaluated in terms of e-log ’, consolidation properties for CSSM can be derived using an
assumed K0 to calculate p’ and adding 1 to void ratio to obtain specific volume, as shown in

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 398

Figure 1(b) and 1(c). The critical state line (CSL) can be projected onto the v-ln p’ space as
shown in Figure 1(c). The slope of NCL or CSL is equal to λ as shown in Figure 1(c). One of the
main limitations of the incremental consolidation test is the 10 to 14 days required to complete
one test. This time can be reduced using constant rate of strain consolidation, if the equipment is
available. However, substantial time and effort is still required.
Burland (1990) proposed the intrinsic compression line (ICL) based on one-dimensional
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

consolidation tests on reconstituted clays with an initial water content between 1.0 to 1.5 LL,
where 𝑤𝐿 indicates the liquid limit. Burland’s method can be used to predict the compression
index (C*c) for vertical effective stresses between 100 and 1000 kPa. The equivalent value of 
can be determined as

Cc* 0.256eL − 0.04


= = = 0.111 LL  Gs − 0.017 (3)
ln (10 ) ln (10 )

where eL is the void ratio of the soil at its liquid limit assuming 100% saturation.
As shown in Figure 1(c), the value of  controls the compression of the soil for normally
consolidated conditions. Because the normally consolidated undrained shear strength is linearly
related to p’, the slope of the v vs. shear strength relationship is also equal to . Wroth and Wood
(1978) used this relationship to estimate  from the Atterberg limits, making assumptions about
the undrained shear strength at the liquid limit (LL) and plastic limit (PL). They also showed that
 can be estimated as

 = Gs A (4)

where A is the slope of the water content versus natural logarithm of undrained shear strength.
This approach allows  to be estimated from a series of undrained shear strength measurements
on normally consolidated or remolded soil at different water contents. The laboratory miniature
vane shear test and the fall cone penetration test are two devices that can be used to rapidly
determine undrained shear strength for this purpose.
Recent research by Farias and Llano-Serna (2016) used the fall cone and miniature vane
shear test to evaluate the deformibility and the strength parameters of kaolin. They found that the
use of the vane shear test produces better constitutive data because it cancels the effect of
anisotropy for the remolded specimen. In addition, they proposed new methods to obtain an
estimation of the critical state strength parameter (M), and these estimations were verified using
undrained triaxial compression tests. They also concluded that fall cone test is a reliable tool to
measure undrained shear strength and deformability of soils if the device is properly calibrated.
However, few, if any, studies have compared values of  estimated from either index tests or
simple undrained shear strength tests to those measured in one-dimensional consolidation test.
The possibility of using simpler, but accurate, means to estimate  provided the motivation for
this paper.

METHODOLOGY

Three different soils were used to compare the methods for determining   Oak Harbor clay,
Memphis silt loam, and Tennessee clay. The index properties for these soils are summarized in

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 399

Table 1. All of the soil specimens were remolded. The effective stress friction angles for the soils
were estimated in the range of 100 to 1000 kPa using Castellanos et al. (2016) and were
subsequently used to estimate K0 for the consolidation tests.

Table 1. Index properties of the soils tested


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Specific Estimated Liquid Plastic Average


USCS
Soil gravity, friction Limit Limit Cone
classification
Gs angle (°)A (%)B (%)C Factor, K
Oak Harbor clay 2.82 31.5 48 22 CL 0.81
Memphis silt loam 2.76 35 36 25 ML 0.78
Tennessee clay 2.73 32 48 21 CL 0.92
Notes: A Castellanos et al. (2016), B Based on fall cone test, C ASTM D4318

For each soil type, conventional one-dimensional consolidation tests were performed on
63.5-mm diameter and 31.75-mm high soil specimens in a fixed-ring type consolidometer
(ASTM D 2435 2011). The soil samples were remolded within the ring at a water content of
about 1.4 times the liquid limit. The vertical effective stresses applied during the test ranged
from 4.8 to 6943 kPa. The load increment ratio was 1.0 during loading and 2.0 during unloading.
The load was sustained sufficiently long to ensure completion of primary consolidation (24 hours
for each load increment step for all tests). The consolidation tests provide the standard
measurement of .
In addition to the consolidation tests, samples of each soil were prepared at water contents
near the liquid limit and tested with both the fall cone test (BS 1377-2:1990) and the miniature
vane shear test (ASTM D4648). The fall cone tests used a cone with a standardized tip angle of
30° and mass of 80 g. The vane shear tests used a 12.7 mm by 12.7 mm vane, which was rotated
by hand. These tests provided measurements of penetration and rotation resistance, respectively,
that can be related to undrained shear strength. The fall cone tests were also used to determine
the liquid limit. Multiple measurements were made at each water content.
In order to determine undrained shear strength at lower water contents, consolidated
specimens of the remolded soils were required. Therefore, the soil samples were remolded in a
custom metal tube with the dimensions of 75 mm diameter and 75 mm height at water contents
from 1.4 LL to 1.00 LL. The specimens were then consolidated to approximately the plastic limit
using the one-dimensional consolidation apparatus. After unloading, the two ends of the
specimen were tested with the miniature vane shear and fall cone tests. The fall cone tests on
stiffer specimens used a fall cone mass of about 400 g.
Undrained shear strength was determined for the miniature vane shear test using the
equations provided in ASTM D4648. For the fall cone test, undrained shear strength can be
estimated as:

KQ
su = (5)
h2

where K is the cone factor, Q is the mass of the cone in grams multiplied by the acceleration due
to gravity (9.81 m/s), and h is the penetration in mm. Historically, the cone factor was assumed
to be constant and equal to 0.85 and 0.29 for cone angles of α = 30° and α = 60°, respectively

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 400

(Wood 1990). However, the cone factor is highly influenced by the chosen method for
performing the fall cone test. The most common calibration method to derive the appropriate K
value is through the combined use of the fall cone test and the miniature vane shear test. In this
case, a large number of K values are generated for each soil (Farias and Llano-Serna 2016;
Nasroulla et al. 2016). In order to derive the proper K value for each soil, the undrained shear
strengths derived from companion miniature vane shear tests were used in Eq. (5) to determine
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the K value for each fall cone test. Finally, the average K value was calculated and used as the K
factor for each soil to calculate undrained shear strengths from the fall cone test. The average K
factors are reported in Table 1.

RESULTS

Figure 2(a) presents one-dimensional compression behavior of the three remolded soil
specimens. Differences in this behavior result from the differences in the particle properties and
size distribution. The compression paths start at different initial specific volumes because of
differences in the initial water contents. However, once vertical stresses of about 100 kPa (ln p’ ≈
4) were reached, the virgin compression line has become approximately log-linear.

Figure 2. One-dimensional consolidation tests: (a) full stress range and (b) vertical stresses
in the range of 100 to 1000 kPa

Figure 2(b) presents the one-dimensional virgin compression line (VCL) for each of the soils
over the range of 100 kPa to 1000 kPa suggested by Burland (1990). The VCLs represent the
intrinsic compression behavior derived from reconstituted samples. The slope of each VCL was
determined and the values are summarized in Table 2. Burland’s (1990) correlations were used
to estimate values of  and the VCL of the soils were plotted on Figure 2(b). The values are also
reported in Table 2.
Moreover, Figure 2(a) presents the swelling or unloading curves from the one-dimensional
consolidation tests. The values of K0 during unloading were estimated for each step using the
overconsolidation ratio (OCR) and the estimated friction angle (’), following Kulhawy and
Mayne (1990):

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 401

K0 = (1 − sin  ') OCRsin ' (6)

The values of K0 were limited to account for passive failure during unloading. K0 was then
used to determine p’ from the vertical effective stress. The slopes of the swelling curves (κ) are
presented in Table 2. The values of / ranged from 9 to 27% for these three soils.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Table 2. Comparison of consolidation properties

VCL slope,  Recompression


Soil Wroth & Wood (1978)
1-D Burland
100 to 1000 κ (1-D) κ/λ (%)
Consol. (1990) Full range
kPa
Oak Harbor
0.147 0.134 0.157 0.087 0.0400 27
Clay
Memphis Silt
0.052 0.094 0.080 0.044 0.0078 15
Loam
Tennessee
0.102 0.128 0.113 0.063 0.0096 9
Clay

A total of 96 vane shear tests and 96 fall cone tests were completed on the three soil types.
Figure 3 presents the variation of the natural logarithm of shear strength with water content for
these soils. The undrained shear strength increases as the water content (and specific volume)
decreases. While these estimations are not quite identical to the critical state line, these
relationships approximate a reasonable value for the λ parameter. The undrained shear strength
for each soil test is calculated using Eq. 5 along with the average K factor for each soil.

Figure 3. Variation of undrained shear strength with water content – (a) miniature vane
shear test and (b) fall cone test

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 402
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Relationship between water content and undrained shear strength for higher
consolidation stresses

DISCUSSION

Using the 1-D consolidation tests as comparison, the values of  from the two simpler
methods are evaluated in Table 3. The results show that Burland method provided good
estimations for the clay soils, Oak Harbor clay and Tennessee clay. The Wroth and Wood
method also worked well for the clays when the full range of water contents was considered to
determine A. However, both of these methods substantially overestimate the  value for the
Memphis silt. The liquid limit and plasticity index of the Memphis silt were at the lower end of
the soils used to create Burland’s (1990) correlation. Moreover, the measured  value for the
Memphis silt was equal to the lowest value in Burland’s data set. Less accuracy is expected for a
soil at the boundary of the correlation.
Overall, the results show that by the decrease of plasticity in soil constituents, the two
methods tend to overestimate the values derived for the λ parameter. In fact, as the percentage of
the sands and silts increase in the soil, the two methods become less reliable for the
determination of the λ value.
Figure 4 examines the relationship of undrained shear strength to water content at water
contents corresponding to higher consolidation stresses. Compared to Figure 3(b), the A
parameters are lower, which reduces the estimates of λ. According to Table 3, by limiting the
range of consolidation from 100 to 1000 kPa, the λ parameter is underestimated by the Wroth
and Wood approach. In addition, by limiting the data to this range, the estimation is better for the
low plasticity silt compared to the clays. This is the opposite conclusion to that found from the
use of the fall cone test results over the full range of water contents.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 403

Table 3. Differences in estimated values of 

Percent difference compared to 1-D consolidation


 based on Fall Cone
Soil  based on LL
(Wroth and Wood 1978)
(Burland 1990)
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Full range 100 to 1000 kPa


Oak Harbor clay -8.9 6.8 -40.8
Memphis silt 80.8 53.8 -15.4
Tennessee clay 25.5 10.8 -38.2

CONCLUSIONS

This study has examined and compared three methods for the determination of the slope of
the virgin compression line, λ. While the use of one-dimensional consolidation tests is the
preferred approach, these tests can be both cost and time-prohibitive. For this reason, the use of
simple alternative methods (i.e., fall cone and miniature vane shear tests) to estimate and define
the virgin compression line is desirable. The following conclusions can be made about the
comparisons presented in this paper:
• For the two clay soils, the Wroth and Wood method provided the best estimate of  when
undrained shear strength data was used from tests at water contents ranging from the
liquid to the plastic limit.
• For the silty soil, the Wroth and Wood method provided the closest estimate of 
provided the undrained strength data at higher water contents was excluded. The silt
exhibited high nonlinearity in the relationship between water content and the natural
logarithm of the undrained shear strength.
• An adequate preliminary estimate of  can be obtained from Burland’s correlation as
long as the index properties of the results fall within the range of those used to develop
the correlation.

ACKNOWLEDGMENTS

The authors acknowledge the generous support of the Center for Energy Systems Research at
Tennessee Technological University for providing funding for this work.

REFERENCES

Alamanis, N., P. Lokkas, T. Chrysanidis, D. Christodoulou, and E. Paschalis. 2021. “Assessment


Principles for the Mechanical Behavior of Clay Soils.” WSEAS Trans. Appl. Theor. Mech,
16: 47–61.
Budhu, M. 2020. Soil mechanics and foundations. Wiley.
Burland, J. B. 1990. “On the compressibility and shear strength of natural clays.” Géotechnique,
40 (3): 329–378.
Dastider, A. G., S. Chatterjee, and P. Basu. 2021. “Advancement in Estimation of Undrained
Shear Strength through Fall Cone Tests.” J. Geotech. Geoenvironmental Eng., 147 (7):
4021047. American Society of Civil Engineers.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 404

Farias, M. M., and M. A. Llano-Serna. 2016. “Simple methodology to obtain critical state
parameters of remolded clays under normally consolidated conditions using the fall-cone
test.” Geotech. Test. J, 39 (5): 1–10.
Guo, Z., Z. Wang, J. Lu, D. Li, D. Liang, X. Xie, and X. Wu. 2022. “Application of Methane
Hydrate Critical State Soil Model on Multistage Triaxial Tests of Methane Hydrate-Bearing
Sediment.” Int. J. Geomech., 22 (7): 6022014. American Society of Civil Engineers.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Kulhawy, F. H., and P. W. Mayne. 1990. Manual on estimating soil properties for foundation
design (No. EPRI-EL-6800). Electric Power Research Inst., Palo Alto, CA (USA); Cornell
Univ., Ithaca, NY (USA). Geotechnical Engineering Group.
Le, T., and D. Airey. 2021. “Mechanical behaviour of a weakly structured soil at low confining
stress.” Géotechnique, 1–15. Thomas Telford Ltd.
Nasroulla, N. A., A. S. A. Rashid, R. Kalatehjari, K. A. Kassim, N. M. Noor, and S.
Horpibulsuk. 2016. “Determination of liquid limit of a low swelling clay using different cone
angles.” Appl. Clay Sci., 132: 748–752. Elsevier.
Roscoe, K., and J. B. Burland. 1968. On the generalized stress-strain behaviour of wet clay.
Wood, D. M. 1990. Soil behaviour and critical state soil mechanics. Cambridge university press.
Wroth, C. P., and D. M. Wood. 1978. “Correlation of index properties with some basic
engineering properties of soils.” Can. Geotech. J., 15 (2).
Zhang, J., S.-C. R. Lo, M. M. Rahman, and J. Yan. 2018. “Characterizing monotonic behavior of
pond ash within critical state approach.” J. Geotech. Geoenvironmental Eng., 144 (1):
4017100. American Society of Civil Engineers.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 405

Geochemical Properties and Atterberg Limits of Low Saline Sand-Clay Mixtures

Uddav Ghimire1; Tejo V. Bheemasetti2; Patrick Kozak3; and Lisa Kunza4


1
Graduate Student, Dept. of Civil and Architectural and Engineering Mechanics, Univ. of
Arizona, Tucson, AZ. Email: uddav@arizona.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Assistant Professor, Dept. of Civil and Architectural and Engineering Mechanics, Univ. of
Arizona, Tucson, AZ. Email: tejo.bheemasetti@arizona.edu
3
Graduate Student, Dept. of Atmospheric and Earth Sciences, South Dakota School of Mines,
Rapid City, SD. Email: patrick.kozak@mines.sdsmt.edu
4
Professor, Dept. of Chemistry, Biology, and Health Sciences, South Dakota School of Mines,
Rapid City, SD. Email: lisa.kunza@sdsmt.edu

ABSTRACT

Increase in soil salinity is one of the major effects of climate change, which has direct and
indirect effects on the chemical, hydraulic, and mechanical behavior of soils. Earlier studies have
demonstrated the change in material properties on soils with low-high saline concentrations;
however, the transition of low to high threshold values that influence the fundamental properties
of the soils is unclear. This study is an attempt to evaluate the threshold of low salt
concentrations on changes to soil properties and develop correlations on the geochemical and
consistency limits of soil. Characterization tests and geochemical tests including particle size,
consistency limits, pH, electrical conductivity (EC), sodium adsorption ratio, and cation
exchange capacity were determined. The amount of Ca, Mg, Na, and K present in the soil was
evaluated using atomic absorption spectrometry (AAS). Each test was performed on the seven
different mixes of varying sodium chloride (NaCl) concentration within a low salinity range to
understand its effect on geochemical parameters and consistency limits. Regression analysis was
performed on the experimental results which showed a linear relationship between electrical
conductivity (EC) and salt concentration whereas liquid limit values were defined by two best
linear fit equations for different salt concentrations. A threshold value was defined for the liquid
limit above which the effect of salt concentration is limited.

INTRODUCTION

Soil salinization is a major effect caused by climate change, with serious consequences for
agricultural production, water quality, biodiversity, and soil erosion, particularly in arid and
semiarid regions (Tanji, 1990; Suarez, 2001). The increase in atmospheric greenhouse gases and
the resulting increase in air temperature and a decrease in relative humidity, as well as intense
rainfall events, are all indicators of climate change that have a significant impact on the rate of
increase in soil salinity (IPCC, 2013; Mukhopadhyay et al., 2020). Additionally, it was reported
that excessive groundwater extraction in arid regions may also result in an increase in the salinity
of both soil and groundwater (Dasgupta et al., 2015; Mukhopadhyay et al., 2020). Soil salinity is
a quantification of the concentration of all soluble salts in soil water, and it is commonly
expressed using the electrical conductivity (EC) of soils. The major soluble salts are the cations
including sodium (Na+), calcium (Ca2+), magnesium (Mg2+), potassium (K+), and the anions
including chloride (Cl−), sulfate (SO4 2−), bicarbonate (HCO3 −), carbonate (CO3 2−), and nitrate

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 406

(NO3 −) (Sumner and Miller,1996). Salt-affected soils can be classified into four categories based
on Electrical Conductivity (EC) and Exchangeable sodium percentage (ESP) as shown in Table
1. ESP is the relative amount of the sodium ion present on the soil surface, expressed as a
percentage of the total Cation Exchange Capacity (CEC). The amount of Ca, Mg, Na, and K ions
present in the soil can be determined using laboratory equipment such as a flame emission
spectrophotometer (FES), an atomic absorption spectrophotometer (AAS), or an inductively
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

coupled plasma (ICP). CEC, ESP, and Soil Adsorption Ratio (SAR) can be calculated by
measuring the cations in the soil.

Table 1. Categories of salt-affected soil based on salinity (US Salinity Laboratory,1954)

EC (ds/m) ESP (%) Soil Classification


<4 <15 Slightly saline
>4 <15 Saline
<4 >15 Sodic
>4 >15 Saline-Sodic
1
ds/m = mmho/cm

The presence of salts in the soils changes the chemical, hydraulic, thermal, and mechanical
properties of soils. Of all, the Atterberg limits are the most basic and critical parameters required
by geotechnical engineers for earthen dam construction and the compaction of soils (Smith et
al.,1984). The Atterberg limits including liquid limit (LL) and plastic limit (PL) are a basic
measure of the critical water contents of the soil that signifies the soil consistency. Atterberg
limits were initially used to distinguish between different types of silts and clays (ASTM D2487,
ASTM 2017), however, various attempts have been made to correlate them with various soil
properties including specific surface area, and cation exchange capacity, mineralogy, swelling
behavior, compressibility, shear strength, and compaction characteristics.
Several authors have investigated the impact of pore fluid salinity on Atterberg limits and
showed that increasing salt concentration either decreases or slightly increases the liquid limit
values. The liquid and plastic limits of soils having a higher percentage of
montmorillonite/smectite clay content decrease significantly with an increase in salt
concentration (Anson and Hawkins, 1998; Di Maio et al., 2004; Tiwari et al., 2005; Yukselen-
Aksoy et al., 2008; Ajalloeian et al., 2013). Contrary to this, the liquid limit of kaolinitic clays
and mixed clay minerals slightly increases or remains constant when treated with different salt
concentrations (Sridharan et al., 2002; Ören et al., 2003). Ying et al., 2020 reported a slight
increase in the liquid limit of soil with high content of non-clay minerals when exposed to higher
water salinity concentration. Based on these observations, it can be understood that the
mineralogical composition and clay content of soils are important factors that govern the
variation of the liquid limit with salinity. Atterberg properties of non-expansive soils were often
examined for higher salinity concentrations. Comprehensive studies have not been performed yet
to understand the chemistry and behavior of soil for low-concentration salts, especially on
mixtures of non-clay minerals and montmorillonite. The low salinity of soil mentioned herein is
below the sodic nature, which includes slightly saline, and saline as mentioned in table 1.
This study aims at evaluating and identifying the low concentration salt effect on the liquid
limit and plastic limit of soils. Soils from Northwestern South Dakota which are frost-

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 407

susceptible, and expansive due to montmorillonite clay mineral were sampled and considered for
this study. This research is a part of a comprehensive study to understand the effect of salt
concentrations on the fundamental behavior of soils and their effects on earthen infrastructure.
This study focuses on the effect of NaCl and aims to present a better understanding of the effect
of low salt concentration on Atterberg limits of the soils with varying NaCl concentrations.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

MATERIALS AND METHODS

Representative samples from the Northwestern part of South Dakota were collected where a
significant increase in saline concentrations are observed. A summary of the basic soil
characterization studies of the test soil with respective ASTM standards is presented in table 2.
The particle size distribution of the sample soil is shown in figure 1 along with their
mineralogical composition in figure 2 evaluated from the X-ray diffraction (XRD) tests.

Table 2. Summary of laboratory test results and physical properties of test soil

ASTM Test
Soil property Test Soil
Designation
Coarse Fraction (%) 89 D422-63
Fine Fraction (%) 11 D422-63
Specific Gravity (Gs) 2.79 D854-00
Liquid Limit, LL (%) 71 D4318
Plastic Limit, PL (%) 32 D4318
Plasticity Index, PI (%) 39 D4318
2
USCS classification Well-graded sand with clay D2487-17
2
USCS-Unified Soil Classification System

The geochemical properties of the test soil including pH, electrical conductivity (EC), Cation
Exchange Capacity (CEC), Sodium Adsorption Ratio (SAR), and Exchangeable Sodium
Percentage (ESP) were determined. The amount of Ca, Mg, Na, and K present in the soil was
evaluated using atomic absorption spectrometry (AAS). The samples were prepared using an
ammonium acetate extraction method (Nathan et al.,2012). The data from AAS were analyzed
and salt cation concentrations along with CEC, SAR, and Exchangeable sodium percentage
(ESP) were calculated which are presented in Table 3. CEC was calculated using Equation (1)
with the sum of exchangeable cations of Ca, Mg, K, and Na (Ross & Ketterings,1995; Sumner et
al., 1998).

CEC = ECa + EMg + EK + ENa (1)

Soil Adsorption ratio (SAR) and Exchangeable Sodium Percentage (ESP) were calculated
using Eq. (2) and Eq. (3) respectively (Ross & Ketterings,1995; Commission,2021).

𝑁𝑎+
SAR = 2+ )+(𝑀𝑔2+ )
(2)
√(𝐶𝑎
2

ESP = [100 (-0.0126 + 0.01475 x SAR)] / [1 + (-0.0126 + 0.01475 x SAR)] (3)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 408
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Particle size distribution plot

Table 3. Summary of geochemical test results of the test soil

Geochemical Properties Control sample


pH 5.4
EC (ds/m) 0.5
Ca (ppm) 7678
Mg (ppm) 1221
Na (ppm) 1736
K (ppm) 607
CEC (meq/100 gm) 58
SAR 1.5
ESP(%) 1

To evaluate the effect of low concentrations of salt on the Atterberg test limits, the sample
soils are mixed with different dosages of sodium chloride (NaCl) powder. The percentage of
sodium in the control soil was measured to be 0.17% of the dry weight of the soil. Sodium
Chloride was added to the test soils at different dosage to make 0.2%, 0.4%,
0.6%,0.8%,1%,1.2% and 1.4% of their dry weight. De-ionized water was added to each mix to
create a uniform paste, which was then oven-dried for 12 hours. This was performed to ensure
the uniform distribution of added NaCl over the soil sample. Then, each sample was subjected to
geochemical tests and Atterberg limit tests (ASTM- D4318) to find the respective properties
using the standard methods mentioned in the earlier section.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 409
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. XRD result, mineralogical composition of test soil

RESULT AND DISCUSSIONS

Geochemical tests

The results of geochemical tests on treated samples are presented in Table 4. With an
increase in the salt concentration from 0.17% to 1.4%, the pH of the soils increased. The EC,
CEC, SAR, and ESP were found to increase with the increase in saline concentrations. Figure 3
presents the relationship between the NaCl and geochemical parameters along with the
regression analysis. The result and correlations were found to be consistent with the observations
and explanations made by (Tiwari et al., 2005). The soils were found to be slightly saline to
saline.

Table 4. Geochemical test results of the treated samples

CEC(meq/100 Soil
Sample pH EC(ds/m) gm) SAR ESP(%) Nature
Control-0.17% 5.4 0.4 58 1.5 1 SS3
0.2% treated 6.12 1.8 59 1.8 1.5 SS
0.4% treated 6.98 2.3 68 3.5 4 SS
0.6% treated 8.05 3.2 76 5.3 6 SS
0.8% treated 9.34 3.4 85 7.1 8 SS
1% treated 9.84 4.1 94 8.8 11 S4
1.2% treated 9.6 4.5 102 10.6 13 S
1.4% treated 10.57 4.9 111 12.4 15 S
SS3 – Slightly Saline; S4- Saline

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 410
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

(c) (d)

Figure 3. Geochemical properties of test soil with salt concentrations: (a) EC; (b) CEC; (c)
SAR; (d) ESP

Atterberg Limits

The test results of the liquid and plastic limits are presented in figures 4 and 5. We observed
an increase in liquid limit values with an increase in salt concentration. In general, the liquid
limit increased up to 0.6% of salt concentration with its maximum value at this point. Up to this
point, the LL value increased by about 3.1%, which can be considered a significant increase.
Above 0.6% of salt concentration, nearly constant values of limit liquid are observed. We
observed no further effect of salt concentration from this point up to 1.4% of salt concentration.
These trends can be described by the multiple linear fit equations mentioned below. Whereas the
plastic limit results did not depict any trends in the data.

4.65 ∗ 𝑥 + 70.62 0.17 ≤ 𝑥 ≤ 0.6


LL =
73.3 0.6 < 𝑥 ≤ 1.4

It is commonly known that electrical forces of attraction and repulsion exist between clay
particles. Several researchers have attempted to gain a better understanding of the nature of these

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 411

forces from the perspective of mechanical behavior (Rosenqvist,1955; Mitchell,1956;


Lambe,1958). The interaction between electrical double layers was found to be the primary
cause of repulsive forces between clay particles whereas, Van der Waal's forces made the most
significant contributions to the attractive force. The diffuse double layer can only be used to
study the behavior of expansive montmorillonites and cannot explain the behavior of non-
swelling clays (Sridharan,1991). So, the LL of non-swelling clays could be defined by the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

geometric arrangement of clay particles (clay fabric) rather than the diffuse double layer theory
(Sridharan and Venkatappa Rao,1975; Sridharan et al.,1988; Sridharan,1991). The clay fabric, as
determined by inter-particle forces and depicted as shearing resistance at inter-particle contacts,
had a significant bearing on the consistency limits, and shear strength characteristics of non-
swelling clays. When pore fluid salinity is increased on non-expansive clays, the inter-particle
attraction gets increased causing an increase in associated particle flocculation. This provides
more void space for fluid entrapment, resulting in higher liquid limits (Sridharan,1991).
However, pH creates a barrier to particle flocculation (Sridharan et al., 1988).

Figure 4. Variation of LL with the concentration of NaCl

Figure 5. Variation of PL with the concentration of NaCl

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 412

In the case of Northwestern South Dakota soil, the liquid limit increased with increasing salt
concentration, due to its fewer clay minerals (Muscovite-18.63%, kaolinite-7.75%, 8.63% of
montmorillonite). The results from the test soil are consistent with the previous studies as the
liquid limit of non-expansive soils either increases or remains almost constant with salt
concentrations (Song et al.,2017; Sridharan et al.,2002; Geertsema and Torrance, 2005; Ying et
al.,2021). As the salinity increased from 0.17% to 0.6%, particle flocculation was increased
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

which required more water in void spaces, in turn, LL was found to be increased. Up to this
point, due to the role of multivalent cations and comparatively low pH values, particle
flocculation was favorable, giving rise to LL values. However, above 0.6% of NaCl
concentration, monovalent sodium ions became dominant and due to higher PH values, particle
flocculation was limited so, almost constant LL values were observed. So, a 0.6% NaCl salt
concentration can be defined as a ‘threshold point’, represented by the vertical blue dashed line
in figure 4.

Figure 6. Variation of Liquid Limit with NaCl concentration

Figure 6 represents the data points from this research study in addition to the literature. It
shows the comparison of low salinity data between the current study and available past studies.
Very few LL values were studied for low salt concentrations in the past, which is insufficient to
analyze its role in the liquid limit values. Also, test results are related to the higher salinity based
up on the past studies. A slight increase or almost constant trend of liquid limit values could be
observed with increasing salinity above 1.4% of dry soil mass.

CONCLUSION

The main objective of this study is to evaluate and understand the effects of saline
concentrations on the geochemical and Atterberg limits of the soils. Laboratory experimental
studies were performed on the test soil collected from Northwestern South Dakota. Based on the
experimental results obtained in this study, it was observed that with an increase in salt
concentration, the Electrical conductivity (EC), Cation exchange Capacity (CEC), Soil

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 413

Adsorption Ratio (SAR), and Exchangeable Sodium Percentage (ESP) values increased linearly
(figure 3). The linear models can be used to interpret the geochemical properties for the soils
with saline concentrations up to 1.4%. The liquid limit increased by 3.1% at the beginning, when
salt concentration increased from 0.17% to 0.6% of the dry weight of the soil, then remained
almost constant with further increase in the soil salinity. For this soil with non-clay and
montmorillonite mixtures, 0.6% of salt concentration acts as a “threshold point”, above which
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

soil salinity does not play a significant role on the liquid limit. More studies shall be performed
on strength and hydraulic parameters as well as combined salinity and freeze-thaw effects to
understand their combined effect on geotechnical parameters.

REFERENCES

Ajalloeian, R., Mansouri, H., and Sadeghpour, A. H. (2013). “Effect of saline water on
geotechnical properties of fine-grained soil”. EDGE, 18, 1419-1435.
Anson, R. W. W., and Hawkins, A. B. (1998). “The effect of calcium ions in pore water on the
residual shear strength of kaolinite and sodium montmorillonite”. Geotechnique, 48(6), 787-
800.
Corporation Commission. (2021). Title 165: Corporation Commission Chapter 10: Oil and Gas
Conservation.
Di Maio, C., Santoli, L., and Schiavone, P. (2004). “Volume change behavior of clays: the
influence of mineral composition, pore fluid composition, and stress state”. Mechanics of
materials, 36(5-6), 435-451.
Geertsema, M., and Torrance, J. K. (2005). “Quick clay from the Mink Creek landslide near
Terrace, British Columbia: geotechnical properties, mineralogy, and geochemistry”.
Canadian Geotechnical Journal, 42(3), 907-918.
Kenney, T. C. (1977). “Residual strengths of mineral mixtures”. Proc., 9th Int. Conf. Soil
Mechanics and Foundation Engineering, 1, 155–160.
Lambe, T. W. (1958). “The structure of compacted clays”. Journal of the Soil Mechanics and
Foundations Division, 84(2), 1654-1.
Mitchell, J. K. (1956). “The fabric of natural clays and its relation to engineering properties”. In
Highway Research Board Proceedings (Vol. 35).
Mukhopadhyay, R., Sarkar, B., Jat, H. S., Sharma, P. C., and Bolan, N. S. (2021). “Soil salinity
under climate change: Challenges for sustainable agriculture and food security”. Journal of
Environmental Management, 280, 111736.
Nathan, M. V., Stecker, J. A., and Sun, U. (2012). Soil testing in Missouri: A guide for
conducting soil tests in Missouri (2012).
Ören, A. H., and Kaya, A. (2003). “Some engineering aspects of homoionized mixed clay
minerals”. Environmental monitoring and assessment, 84(1), 85-98.
Regional Salinity Laboratory (US). (1954). Diagnosis and improvement of saline and alkali soils
(No. 60). US Department of Agriculture.
Rosenqvist, I. T. (1955). “Investigations in the clay-electrolyte-water system”. Norwegian
Geotechnical Institute.
Ross, D. S., and Ketterings, Q. (1995). “Recommended methods for determining soil cation
exchange capacity”. Recommended soil testing procedures for the northeastern United
States, 493(101), 62.
Smith, C. W., Hadas, A., Dan, J., and Koyumdjisky, H. (1985). “Shrinkage and Atterberg limits
in relation to other properties of principal soil types in Israel”. Geoderma, 35(1), 47-65.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 414

Song, M. M., Zeng, L. L., and Hong, Z. S. (2017). “Pore fluid salinity effects on
physicochemical-compressive behaviour of reconstituted marine clays”. Applied Clay
Science, 146, 270-277.
Sridharan, A., and Venkatappa Rao, G. (1975). “Mechanisms controlling the liquid limit of
clays”. In Proceedings of the Istanbul Conference on Soil Mechanics and Foundation
Engineering, Vol. 1, pp. 65–74.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Sridharan, A., Rao, S. M., and Murthy, N. S. (1988). “Liquid limit of kaolinitic soils”.
Geotechnique, 38(2), 191-198.
Sridharan, A. (1991). “Engineering behavior of fine-grained soils: a fundamental approach”.
Indian Geotechnical Journal, 21(1), 1-136.
Sridharan, A., El-Shafei, A., and Miura, N. (2002). “Mechanisms controlling the undrained
strength behavior of remolded Ariake marine clays”. Marine Georesources and
Geotechnology, 20(1), 21-50.
Suarez, D. L. (2001). “Sodic soil reclamation: Modelling and field study”. Soil Research, 39(6),
1225-1246.
Sumner, M. E., and Miller, W. P. (1996). “Cation exchange capacity and exchange coefficients”.
Methods of soil analysis: Part 3 Chemical methods, 5, 1201-1229.
Tanji, K. K. (1990). Agricultural salinity assessment and management. In American Society of
Civil Engineers (Vol. 54, pp. 413-415).
Tiwari, B., Tuladhar, G. R., and Marui, H. (2005). “Variation in residual shear strength of the
soil with the salinity of pore fluid”. Journal of Geotechnical and Geoenvironmental
engineering, 131(12), 1445-1456.
Ying, Z., Cui, Y. J., Duc, M., Benahmed, N., Bessaies-Bey, H., and Chen, B. (2021). “Salinity
effect on the liquid limit of soils”. Acta Geotechnica, 16(4), 1101-1111.
Yukselen-Aksoy, Y., Kaya, A., and Ören, A. H. (2008). “Seawater effect on consistency limits
and compressibility characteristics of clays”. Engineering Geology, 102(1-2), 54-61.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 415

Effect of Grain Size of Granular Soils on Shear Wave Velocity and Electrical Resistivity for
Levee Health Monitoring

Brittany M. Russo1; Adda Athanasopoulos-Zekkos, Ph.D., M.ASCE2;


and Jongchan Kim, Ph.D., Aff.M.ASCE3
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
GeoSystems Engineering, Dept. of Civil and Environmental Engineering, Univ. of California,
Berkeley, CA. Email: bmrusso@berkeley.edu
2
GeoSystems Engineering, Dept. of Civil and Environmental Engineering, Univ. of California,
Berkeley, CA. Email: adda.zekkos@berkeley.edu
3
GeoSystems Engineering, Dept. of Civil and Environmental Engineering, Univ. of California,
Berkeley, CA. Email: jkim3139@berkeley.edu

ABSTRACT

Levees are geologically complex earthen structures that vary laterally and vertically. Current
levee inspection practices consist of mainly visual inspections of the surface with limited
instrumentation that measures data at discrete locations. To better characterize and monitor the
subsurface of these highly complex and spatially distributed systems, non-invasive geophysical
methods, such as the multichannel analysis of surface waves (MASW) and electromagnetic
induction (EMI), can be used to map the geophysical properties, i.e., shear wave velocity and
apparent electric resistivity, respectively. This study focuses on investigating the effect of soil
grain size on shear wave velocity and electric resistivity measurements conducted in the
laboratory for a range of relative density, water content, and confining stress values. The testing
program involved two types of sand: Ottawa C109 sand and Nevada sand, which have different
grain sizes, with a D50 of 0.36 mm and 0.18 mm, respectively. It is shown that both shear wave
velocity and electric resistivity measurements were affected by grain size and can therefore be
used to distinguish between different types of sands at depth. Laboratory testing showed that the
coarser sand has higher shear wave velocity at lower water contents and that shear wave velocity
decreases with increasing water content. Finer sand particles exhibited lower electrical resistivity
for all soil densities and water contents compared to the coarser sand specimens, but the
relationship was more pronounced at lower water contents.

INTRODUCTION

Levees are spatially extensive critical infrastructure used for flood management. They can be
either man-made or formed through natural fluvial processes which results in complicated
lithologies in the subsurface, resulting in many potential internal failure mechanisms. These
systems were deemed critical enough to be included in the ASCE Infrastructure Report Card.
Levees have repeatedly been given a grade of “D” in the report card indicating that these systems
are in poor health. This means that these systems have significant deterioration which puts them
at a strong risk for failure (ASCE 2019). With the increasing demand on these systems due to
climate change, the health of levees need to be improved, and one way to do this is through more
standardized inspection practices.
Most levee inspections in the United States are done visually looking for surface evidence of
failure (USSD 2016). Visual inspection is an important component of inspection, but many

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 416

failure mechanisms of levees are internal, and by the time evidence of those failure mechanisms
are visible at the surface, it may be too late for action. Some instrumentation, such as
piezometers, extensometers, and moisture content sensors has been used for subsurface
inspection practices, but these measure a single response at a single location (USSD 2016). The
single data points measured by these systems are not spatially distributed enough for levees
which can span hundreds of miles in length.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Using subsurface imaging techniques such as geophysics-based approaches can provide a


more non-invasive continuous characterization of the subsurface, a missing component from
current inspection practices, and provide a more complete picture of subsurface health. Two non-
invasive geophysical methods that can be used for subsurface characterization of levees are: 1.)
Multichannel of Surface Waves (MASW) and 2.) Electromagnetic Induction (EMI). MASW is a
surface waves seismic method that can be used to determine the shear wave velocity (Vs) profile
along an array of sensors. Shear wave velocity can then be correlated to soil stiffness, density, as
well as liquefaction vulnerability (Inazaki and Sakamoto 2005; Park et al. 1999). EMI uses
electromagnetic waves to infer the apparent resistivity (ρ) of the subsurface which can be used as
a proxy for soil moisture (Asch et al. 2008; Bishop et al. 2003; Tresoldi et al. 2018).
The laboratory data collected on Vs and ρ serves the purpose of supporting interpretation of
field measurements of Vs and ρ profiles from MASW and EMI. A range of different geological
and environmental factors are being tested in the laboratory to build a library of Vs and ρ values
under a range of water contents (w), relative density (Dr), and confinement effects that can be
used to understand the field data more accurately. This paper focuses on one factor being
explored in the laboratory, the effect of grain size on Vs and ρ. Changes in grain size can be
easily characterized by invasive sampling techniques, but the laboratory data will support non-
invasive imaging techniques for more continuous characterization of the subsurface. In this
study, two sands of different grain size are tested under different Dr, w, and confining pressures.
Sands are being tested in this setup because borehole logs of the levees where field testing is
being conducted show that the subsurface is comprised of sands and silty sands (Hultgren-Tillis
Engineers 2021).
A common method for measuring Vs in the lab is by using bender elements. Park and Lee
(2013) used bender elements to generate elastic waves to understand how freezing conditions
affected the characteristics of sand-silt mixtures while Aris et al. (2012) were investigating how
different modes of soil deposition affected the structural anisotropy of soils during undrained
shear. A lot of work has been done to understand large and small-strain stiffness in sands using
bender elements under different conditions such as at critical state (Robertson et al. 1995) and
with varying amounts of plastic and non-plastic fines (Carraro et al. 2009). Hardin and Richart
(1963) extensively explored how different factors affect Vs, such as confining pressure, void
ratio, w, and soil grain characteristics.
Measuring ρ in a laboratory setting is not as extensively researched as Vs and has mainly
focused on how w affects ρ. Pandey et al. (2015) explored how a quarried sandy soil from Perth
is affected by w and Dr. Kang and Lee (2015) used the ρ of the soil to understand the freezing-
thawing cycle of silty sands. Kibria and Hossain (2012) was interested in how the w and unit
weight (γ) of a clayey soil affected ρ while Tang et al. (2018) used the resistivity of clays to
understand evaporation and cracking. Electrical resistivity of soils has also been used to develop
models that can predict hydraulic conductivity or the degree of saturation of a soil sample (Won
et al. 2019; Cardoso and Dias 2017).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 417

EXPERIMENTAL SETUP

Materials and Preparation

To look at the effect of grain size on Vs and ρ, two sub-rounded silica sands were used with
similar gradations, shown in Figure 1. The coarser sand used was Ottawa C109 which had a
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Coefficient of Uniformity (Cu) of 1.60, a D50 of 0.36 mm, and a Coefficient of Curvature (Cc) of
0.90. The finer grained sand used in the study was Nevada sand with a Cu of 2.0, D50 of 0.18
mm, and a Cc of 1.13. The minimum and maximum γ for each soil was determined in accordance
with the Japanese Standard JIS A 1224:2009, resulting in a minimum γ of 14.8 kN/m3 for C109
and 14.4 kN/m3 for Nevada. The maximum γ for C109 was 17.3 kN/m3 while Nevada Sand was
17.2 kN/m3. For C109, the minimum and maximum void ratio is 0.50 and 0.76, respectively. For
Nevada, the minimum and maximum void ratio is 0.51 and 0.81, respectively.

Figure 1. Gradation Curve Comparing Medium Coarse Sand (C109) and Fine Sand
(Nevada Sand)

Specimen preparation was the same for all specimens. Each soil sample was first oven dried
and then deionized water was mixed into the sample until the goal w was reached. Three w were
tested, 1%, 5%, and 10%. For the C109 sand, the degree of saturations ranged between 3.8% to
45.7% while the degree of saturations for the Nevada sand ranged between 3.7% to 44.3%. After
the soil was at the desired w, each sample was compacted to a specified Dr using three equal lifts
at 30%, 50%, and 70%.
To look at the effect of confinement, each test measured Vs and ρ at seven stresses: 0 kPa, 5
kPa, 10 kPa, 25 kPa, 50 kPa, 75 kPa, and 100 kPa. This stress range was chosen based upon the
depth of investigation of the field equipment for subsurface electrical resistivity measurements.

Shear Wave Velocity

Bender elements were used to measure the Vs of the soil. Bender elements are piezoelectric
transducers imbedded into a soil sample that bend with an applied voltage which generates a

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 418

wave that travels through the soil. Two bender elements are needed to measure Vs, as shown in
the soil cell schematic in Figure 2. One bender element acts as a transceiver and is excited
through an applied voltage which generates the wave. The second bender elements act as a
receiver and bends when the wave interacts with the sensor, generating a waveform (Viggiani
and Atkinson 1995).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. Schematic of Cell Design used for Testing Shear Wave Velocity and Electrical
Resistivity in the Laboratory

The bender elements used in this study were cut to 8 mm by 4 mm by 0.5 mm from T220-
A4BR-2513YE piezoelectric material. Coaxial cables were soldered in a parallel circuit to the
piezoelectric ceramic. Layers of epoxy and M-coat A polyurethane were applied to all sides of
the bender elements for water proofing A layer of conductive silver paint was applied to improve
grounding followed by another layer of epoxy.
An Agilent 33500B Series waveform generator was connected to the transceiver bender
element to create an input wave using a peak-to-peak voltage of 10 V while a Tektronix MSO
2012 Mixed Signal Oscilloscope was connected to the receiver bender element to measure the
output wave. A frequency sweep with a range between 1 kHz to 25 kHz was performed for each
soil before testing to determine the resonant frequency, which is the frequency with the largest
amount of energy transmitted (Montoya et al. 2012). The frequency with the largest peak
magnitude was established as the resonant frequency and was used as the frequency for testing
for that soil.
The Vs of the soil was calculated using:
𝐿
𝑉𝑠 = ∆𝑡 (1)

Where Vs was measured in m/s, L is the tip-to-tip distance between the two bender elements
in meters and Δt is the travel time difference between the input and received wave in seconds.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 419

The difference in travel time was determined by finding the difference in time between the first
arrival of the input wave and the first arrival of the received wave.

Electrical Resistivity

Soil ρ was measured using four stainless steel rods that were cut to 16 mm in length with a 2
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

mm diameter. Two rods were installed in the top cap and two rods were installed in the bottom
cap following the modified-four-electrode method, as shown in Figure 2. Two rods were used as
current electrodes that allow for the flow of current through the soil and the other two rods were
potential electrodes that measure the potential difference. The resistance of the soil was
measured using a GW Instek 8300 Precision LCR meter with an input voltage of 1 V and an
operating frequency of 1,000 Hz which was determined through a frequency sweep of three
different NaCl solutions of known resistivity (Yun et al. 2016; Cho et al. 2004; Kang and Lee
2015).
The resistance measured by the LCR meter can be converted to resistivity through a
geometric factor:

ρ = kR (2)

Where ρ was measured in Ohm-meters (Ωm), R is the resistance of the soil (Ω), and k is the
geometric factor in meters. The k is function of the cell geometry, the electrodes, and the
thickness of the soil sample and is determined experimentally by measuring the resistance of
many NaCl solutions with a known ρ. Then k can be back calculated using Equation (2)
(Yamashita 1987).

RESULTS AND DISCUSSION

Nine tests were performed on each sand with varying w of 1%, 5%, and 10% and Dr of 30%,
50%, and 70%. The results showing the effect of different grain size on Vs and ρ are shown in
Figures 3 and 4, respectively.
Figures 3a, 3d, and 3g show the Vs comparison of samples with w of 1% while Figures 3c,
3f, and 3i show the Vs comparison of samples with w of 10%. The data shows that there is an
overall trend of decreasing Vs with increasing w for all the Dr, which is consistent with work
done by Yang et al. (2008) and Dong and Lu (2016). In unsaturated condition, a higher capillary
force is expected as w decreases. This higher capillary pressure (Pc) can attribute the higher Vs
since Vs is mainly affected by effective stress of soils. Capillary pressure can be defined by Kim
et al. (2021) and specific surface (Ss) can be estimated by Santamarina et al. (2002):
𝑇𝑠 𝑆𝑠 𝜌𝑚
Pc = 2 (3)
𝑒

3(𝐶𝑢 +7)
Ss = 4𝜌 (4)
𝑤 𝐺𝑠 𝐷50

where Ts is the interfacial tension (N/m) and assumed to be 0.072 N/m, Ss is measured in m2/g,
and ρm is the mineral density (kg/m3) which was assumed to be 2650 kg/m3. The Ss was
calculated to be 0.014 m2/g for Nevada sand and 0.007 m2/g for C109 sand using Equation (4).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 420

The first-order estimated Pc for the tests conducted at w =1% range from 7.56 kPa to 9.01 kPa
for Nevada sand and 3.79 kPa to 4.44 kPa for C109 sand, depending on the packing density.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Results showing shear wave velocity as a function of confinement between a


medium sand and a fine sand (a.) w = 1% and Dr = 30%, (b.) w = 5% and Dr = 30%, (c.) w
= 10% and Dr = 30%, (d.) w = 1% and Dr = 50%, (e.) w = 5% and Dr = 50%, (f.) w = 10%
and Dr = 50%, (g.) w = 1% and Dr = 70%, (h.) w = 5% and Dr = 70%, (i.) w = 10% and Dr
= 70%.

Even though the Nevada sand has a higher Pc compared to C109 sand, the difference in Pc
between the two soils is small. Because the difference is negligible, it can be concluded that the
difference in Vs between C109 and Nevada sand is attributed to particle size rather than Pc.
More work needs to be done in the future to develop soil water characteristic curves (SWCC) to
better understand Pc at the varying w for these tests.
From Figure 3, there is relatively minor difference in Vs between grain size at higher w. We
can assume that the bulk density of the material is going to increase with increasing w. Since Vs
is a function of the shear modulus and the bulk density of the material, if the bulk density
increases but the shear modulus remains constant (since water has no shear resistance), then the
Vs must decrease, which is observed in the data.
Based upon the results from this data, Vs can be used to distinguish between sand grain sized
particles in field conditions with low moisture but becomes more difficult to distinguish in soils
with higher moisture.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 421

For all w and all Dr, the larger particle size has a greater ρ compared to the smaller particle
size. While this relationship holds true for all cases that were tested, this relationship is more
prominent in soil samples that have lower w (Figures 4a, 4d, and 4g). With lower w, there is a
more significant difference in ρ between particle size. As the w of the soil sample increases, the
difference in ρ becomes smaller and harder to distinguish between the particle size. This is
expected as w significantly affects the ρ of the soil, which is clearly represented by the data. With
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

lower w, the ρ are higher and as w increases, the ρ of the soils significantly decrease.

Figure 4. Results showing electrical resistivity as a function of confinement between a


medium sand and a fine sand (a.) w = 1% and Dr = 30%, (b.) w = 5% and Dr = 30%, (c.) w
= 10% and Dr = 30%, (d.) w = 1% and Dr = 50%, (e.) w = 5% and Dr = 50%, (f.) w = 10%
and Dr = 50%, (g.) w = 1% and Dr = 70%, (h.) w = 5% and Dr = 70%, (i.) w = 10% and Dr
= 70%.

According to the data, the Dr of the soil has relatively less effect on the ρ compared to the w.
The loose soil samples (Figures 4a, 4b, and 4c) have similar ranges of ρ values compared to the
denser soil samples (Figures 4g, 4h, and 4i). As a result, the effect of grain size for ρ is
independent of Dr. Due to the clear relationships shown in the laboratory data between ρ, grain
size, and w, using field methods that measure ρ of the subsurface can be used to distinguish
between grain size and locations with differing w.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 422

CONCLUSION

The data from the laboratory testing shows that it is possible to distinguish between relatively
finer and coarser grained sand particles using Vs at low w. As the w increases, there is a decrease
in Vs due to increasing bulk density of the sample. At higher w, it becomes difficult to
distinguish between coarse- and fine-grained sand particles because of the minor Vs differences
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

in soils. The laboratory data has also shown that ρ can be used to distinguish between particle
size. Coarser grained sand particles have a higher ρ than finer grained sand particles for all Dr
and w.
By understanding the relationships between Vs and ρ for grain size at a range of w and Dr in
the laboratory, better interpretation of non-invasive field methods can be conducted for levee
health monitoring. At low w, if the Vs subsurface profile shows an area with an increase in Vs
while the ρ profile shows an increase in the same location, it can be assumed that a coarser grain
sand could be present potentially without the need for invasive soil sampling. For areas with
higher w, a coarse-grained soil can be detected compared to the surrounding soil if there is no
change or a slight decrease in Vs combined with a decrease in the ρ profile at the same location.
More work needs to be done by testing Vs and ρ with higher w to match field w measurements of
levees at the test sites.

REFERENCES

Aris, M., Benahmed, N., and Bonelli, S. (2012). A laboratory study on the behavior of granular
material using bender elements. European Journal of Environmental and Civil Engineering.
ASCE. (2019). Report Card for California’s Infrastructure. ASCE. Retrieved from
INFRASTRUCTUREREPORTCARD.ORG/CALIFORNIA.
Asch, T. H., Deszcz-Pan, M., Burton, B. L., and Ball, L. B. (2008). Geophysical characterization
of the American River levees, Sacramento, California, using electromagnetics, capacitively
coupled resistivity, and DC resistivity. Virginia: USGS.
Bishop, M., Dunbar, J., and Peyman-Dove, L. (2003). Integration of remote sensing (LIDAR,
electromagnetic conductivity) and geologic data toward the condition assessment of levee
systems. SPIE, 4886, pp. 400-407.
Cardoso, R., and Dias, A. S. (2017). Study of the electrical resistivity of compacted kaolin based
on water potential. Engineering Geology, 226, 1-11.
Carraro, J. A., Prezzi, M., and Salgado, R. (2009). Shear Strength and Stiffness of Sands
Containing Plastic or Nonplastic Fines. Journal of Geotechnical and Geoenvironmental
Engineering, 135(9), 1167-1178.
Cho, G., Lee, J.-S., and Santamarina, J. C. (2004). Spatial Variability in Soils: High Resolution
Assessment with Electrical Needle Probe. Journal of Geotechnical and Geoenvironmental
Engineering, 843-850.
Dong, Y., and Lu, N. (2016). Dependencies of Shear Wave Velocity and Shear Modulus on Soil
Saturation. Journal of Engineering Mechanics, 142(11), 04016083.
Hardin, B. O., and Richart, F. E. (1963). Elastic Wave Velocities in Granular Soils. Journal of
the Soil Mechanics and Foundations Division, 89(1), 33-65.
Hultgren-Tillis Engineers. (2021). Preliminary Geotechnical Investigation, Three Seepage Sites,
Reclamation District 3, Grand Island, Sacramento County, California. Concord: Hultgren-
Tillis Engineers.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 423

Inazaki, T., and Sakamoto, T. (2005). Geotechnical Characterization of Levee By Intergrated


Geophysical Surveying.
Kang, M., and Lee, J. S. (2015). Evaluation of the freezing-thawing effects in sand-silt mixtures
using elastic waves and electrical resistivity. Cold Regions Science and Technology, 113, 1-
11.
Kibria, G., and Hossain, M. S. (2012). Investigation of Geotechnical Parameters Affecting
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Electrical Resistivity of Compacted Clays. Journal of Geotechnical and Geoenvironmental


Engineering, 138(12), 1520-1529.
Kim, J., Won, J., and Park, J. (2021). Effects of water saturation and distribution on small-strain
stiffness. Journal of Applied Geophysics, 186, 104278.
Montoya, B. M., Gerhard, R., DeJong, J. T., Wilson, D. W., Weil, M. H., Martinez, B. C., and
Pederson, L. (2012). Fabrication, Operation, and Health Monitoring of Bender Elements for
Aggressive Environments. Geotechnical Testing Journal, 35(5), 728-742.
Pandey, L. M., Shukla, S. K., and Habibi, D. (2015). Electrical resistivity of sandy soil.
Geotechnique Letters, 5, 178-185.
Park, C.-B., Miller, R.-D., and Xia, J. (1999). Multichannel analysis of surface waves.
Geophysics, 63(3), 800-808.
Park, J.-H., and Lee, J.-S. (2014). Characteristics of elastic waves in sand-silt mixtures due to
freezing. Cold Regions Science and Technology, 99, 1-11.
Robertson, P. K., Sasitharan, S., Cunning, J. C., and Sego, D. C. (1995). Shear-Wave Velocity to
Evaluate In-Situ State of Ottawa Sand. Journal of Geotechnical Engineering, 121(3), 262-
273.
Santamarina, J. C., Klein, K. A., Wang, Y. H., and Prencke, E. (2002). Specific surface:
determination and relevance. Canadian Geotechnical Journal, 39, 233-241.
Tang, C.-S., Wang, D.-Y., Zhu, C., Zhou, Q.-Y., Xu, S.-K., and Shi, B. (2018). Characterizing
drying-induced clayey soil desiccation cracking process using electrical resistivity method.
Applied Clay Science, 152, 101-112.
Tresoldi, G., Arosio, D., Hojat, A., Longoni, L., Papini, M., and Zanzi, L. (2018). Tech-Levee-
Watch: experimenting an integrated geophysical system for stability assessment of levees.
Rendiconti Online Societa Geologica Italiana, 46, 38-43.
USSD (United States Society on Dams). (2016). Monitoring Levees. Denver: United States
Society on Dams.
Viggiani, G., and Atkinson, J. H. (1995). Interpretation of bender element tests. Geotechnique,
45(1), 149-154.
Won, J., Park, J., Choo, H., and Burns, S. (2019). Estimation of saturated hydraulic conductivity
of coarse-grained soils using particle shape and electrical resistivity. Journal of Applied
Geophysics, 167, 19-25.
Yamashita, M. (1987). Resistivity correction factor for the four-probe method. Journal of
Physics E: Scientific Instruments, 20, 1454-1456.
Yang, S., Lin, H., Kung, J., and Liao, J. (2008). Shear wave velocity and suction of unsaturated
soil using bender element and filter paper method. Journal of GeoEngineering, 3(2), 67-74.
Yun, T. S., Narsilio, G. A., Santamarina, J. C., and Ruppel, C. (2006). Instrumented pressure
testing chamber for characterizing sediment cores recovered at in situ hydrostatic pressure.
Marine Geology, 229, 285-293.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 424

Deep Learning-Based Segmentation for Field Evaluation of Riprap and Large-Sized


Aggregates

Jiayi Luo1; Haohang Huang2; Issam I. A. Qamhia3; John M. Hart4; and Erol Tutumluer5
1
Newmark Civil Engineering Laboratory, Univ. of Illinois at Urbana-Champaign, Urbana, IL
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Newmark Civil Engineering Laboratory, Univ. of Illinois at Urbana-Champaign, Urbana, IL
3
Newmark Civil Engineering Laboratory, Univ. of Illinois at Urbana-Champaign, Urbana, IL
4
Computer Vision and Robotics Laboratory, Univ. of Illinois at Urbana-Champaign, Urbana, IL
5
Newmark Civil Engineering Laboratory, Univ. of Illinois at Urbana-Champaign, Urbana, IL
(corresponding author). Email: tutumlue@illinois.edu

ABSTRACT

Aggregate materials play essential roles in the sustainable performance of engineering


applications. Size and shape properties of these aggregates determine the composition and
packing of the aggregate assemblies, which impact layer strength, modulus, and deformation
response under wheel loading. Current state-of-the-practice methods to evaluate aggregate
properties rely on visual inspection, manual measurements, material sampling, and laboratory
sieve analysis, as well as ground penetrating radar (GPR) for on-site inspection. Advanced
aggregate imaging systems developed to date primarily focused on measurement of separated or
slightly contacting aggregate particles, which is not applicable for densely stacked aggregate
assemblies under field conditions. In this regard, development of a reliable, accurate, and cost-
effective field evaluation technique is crucial. This paper presents an innovative approach using
deep learning-based image segmentation techniques for evaluating aggregate materials in the
field. The approach requires preparing an image data set comprised of aggregate images
collected from various field and laboratory sites, where each image is manually labeled to
identify individual particle locations by tracing their boundaries. Then, a state-of-the-art object
detection and segmentation framework, called Mask R-CNN, is trained on the task-specific data
set to develop the image segmentation kernel. Finally, the trained segmentation kernel is used to
segment individual particles, and post-processing provides the morphological indices for the
segmented particles. This segmentation approach has been utilized for the field evaluation of
large-sized riprap materials and medium-sized ballast materials, with further performance tuning
to address various field conditions. The segmentation results show good agreement with the
ground-truth labeling and greatly improves the accuracy and efficiency of aggregate field
evaluation procedures. Based on the presented findings, the deep learning-based segmentation
approach can serve as an efficient and reliable tool for determining aggregate shape and
morphological properties in field evaluation applications.

INTRODUCTION AND BACKGROUND

Particle size and morphological properties of aggregate materials are among the primary
factors that determine the performance and characteristics of the aggregate skeleton in various
structures such as Portland cement concrete layers (Fowler and Quiroga 2003), railway ballast
(Wnek et al. 2013), unbound or bound pavement layers (Tutumluer and Pan, 2008), and riprap
materials for slope stability and erosion control (Lagasse 2006). Obtaining the morphological

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 425

properties helps to understand the behavior of an aggregate material (Barrett 1980). Current
state-of-the-practice aggregate evaluation methods mainly rely on visual inspection, manual
measurements, field sampling and sieve analysis. The visual inspection and manual
measurements are subjective and depend on the level of practitioner’s experience. Further,
obtaining a representative aggregate field sample can be challenging since obtaining undisturbed
aggregate samples are not feasible due to the nature of unbound aggregates (Stark and Wilk
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2018). In this regard, reliable, accurate and cost-effective aggregate analysis approaches are
essential for field aggregate evaluation.
Over the past few decades, deep learning-based image analysis methods have gained
increasing attention due to their fast adaptation and high accuracy. A variety of aggregate
imaging systems have been proposed and developed (Huang et al. 2020). Imaging-based
aggregate analysis consists of an image segmentation module followed by a post-processing
module to calculate the task-specific deliverables. Image segmentation captures and extracts the
region of target particle(s) from the image background, which is a critical step for informative
image acquisition and feature extraction. Existing aggregate imaging systems were proposed and
designed for analyzing aggregate materials under different arrangements (separated or densely
stacked) and conditions (laboratory or field conditions). Masad et al. (2007) developed, and
Gates et al. (2011) improved the prototype of a laboratory imaging system, Aggregate Imaging
System (AIMS), which analyzes multiple aggregates manually separated and placed. Tutumluer
et al. (2000) developed and Moaveni et al. (2013) improved a laboratory imaging system, the
Enhanced-University of Illinois Aggregate Image Analyzer (E-UIAIA), which captures
individual aggregates sequentially from three orthogonal views. AIMS and E-UIAIA imaging
systems both provide an efficient analysis for separated aggregates under laboratory conditions.
For field conditions where the random arrangement of aggregates and complicated
environmental factors often bring challenges to image analysis, Huang et al. (2020) recently
developed a deep learning-based field imaging system for aggregate materials.

OBJECTIVE AND METHODOLOGY

The objective of this paper is to introduce a deep learning-based approach for the field
evaluation of aggregate materials. The proposed approach consists of developing an image
segmentation module which generates the intermediate segmented images for post-processing,
and a post-processing module which calculates task-specific deliverables by evaluating the
segmented aggregate particles. The overall development method can be abstracted by a flowchart
delineating the preparation – training – analysis pipeline as shown in Figure 1.
The methodology adopted to evaluate the aggregate material entails the following three
aspects:
• Due to the inherent data-driven property of deep learning techniques, the performance of
the image segmentation module is heavily dependent on the quality of the image
database, which guides how the model obtains the required knowledge and learns to
extract the underlying feature maps from the given aggregate images. Therefore, a
sufficient number of aggregate images including large-sized riprap and medium-sized
ballast materials from various sources are collected and manually labeled during the
preparation step to establish a high-quality database.
• During the training step, a state-of-the-art deep learning framework for image
segmentation, Mask Region-based Convolutional Neural Network (Mask R-CNN) (He et

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 426

al. 2017), is selected as the architecture for the segmentation kernel. After training with
the established database, the segmentation is expected to detect and segment out the
aggregate particles under various environmental conditions. Meanwhile, several model
hyper-parameters are tested and fine-tuned to achieve optimal performance of the final
segmentation kernel during training.
• As the final step, task-specific deliverables are calculated using the segmented particle
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

information. For large-sized riprap materials, the size and shape properties as well as
gradation are calculated with reference to a calibration object. Then the collective
statistical analysis of the particle properties in the aggregate image can be illustrated in
the form of a cumulative distribution. For medium-sized ballast material, an index for
each image, i.e., Percent of Degraded Segments (PDS), which is defined as the
percentage of the total area of particle segments compared to the total area of the image is
calculated (Tutumluer et al. 2017). For ballast images, a statistical regression analysis is
then performed to establish a comprehensive ballast degradation model relating the
commonly used Fouling Index (FI) (Selig and Waters 1994) to the PDS value.

Figure 1. Flowchart of deep learning-based image segmentation and aggregate evaluation


approach.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 427

AGGREGATE IMAGE DATABASE PREPARATION

The established aggregate database is task-specific and should contain the necessary
information to guide the segmentation kernel to accomplish the segmentation task. The aggregate
images in the dataset are collected following this criteria: (a) the database should include
aggregate images from various field as well as laboratory conditions, (b) the database should
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

cover aggregate materials with different sizes and colors, and (c) the database should contain
aggregate images with varying placement of forms and lighting conditions. The source
information and descriptions of the aggregate image database are summarized in Table 1.
To provide the deep learning model with ground-truth data during the training process, the
manual segmentation step, known as “image labeling,” is crucial to indicate where the human
eye detects and identifies the locations and regions of all the particles present in each aggregate
image. Each aggregate particle is marked using a high-degree polygon with all vertex
coordinates recorded. The VGG Image Annotator (VIA) (Dutta et al. 2016), an open-source
image annotation software, is selected as the auxiliary tool to facilitate the labeling task. The
labeling process is guided by the following criteria: (a) each visually distinguishable particle
should be carefully labeled, (b) the polygonal lines should approximate the boundaries with
small deviations from the actual shape, and (c) incomplete and occluded particles near the image
boundaries should be labeled so that the segmentation model can obtain consistent performance
at different locations from an image. Examples of raw and labeled images are given in Figure 2.

Table 1. Source Information and Description of Aggregate Image Database

Number of
Image Source Location
Images
6 Aggregate producer - Ocoya, Illinois
14 Aggregate producer - Milan, Illinois
Field aggregate images collected from Aggregate producer - Hillsdale,
different quarry producers in Illinois 100
Illinois
Aggregate producer - Kankakee,
44
Illinois
18 TTCI HTL track in Pueblo, Colorado
BNSF track near Kansas City,
14
Field aggregate images collected from Missouri
different railroad sites
2 UP track in Martinton, Illinois
24 CN track in Champaign, Illinois
Laboratory aggregate images collected Illinois Center for Transportation
186
from lab-engineered samples (ICT) - UIUC

Following the procedure and criteria established above, 408 aggregate images were carefully
labeled and utilized to establish the aggregate image database for training. This labeled database
serves as the human vision-based ground-truth for the deep learning framework described in
detail in the following section.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 428
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Figure 2. Example aggregate images: (a) before labeling, and (b) after labeling.

AUTOMATED IMAGE SEGMENTATION MODEL SELECTION

The key component of the algorithm pipeline is the segmentation kernel, which directly
impacts the evaluation accuracy of the final outputs. This segmentation kernel is essentially a
specialized image segmentation model for aggregate application, after being trained with the
established aggregate database. When receiving a new aggregate image, the segmentation kernel
identifies and extracts each of the individual particles in a user-independent manner.
Considering the requirement to explicitly detect and segment out every instance of the
particles located in an image, the state-of-the-art instance segmentation framework, Mask R-
CNN, is selected as the architecture of the segmentation kernel. The Mask R-CNN is a general-
purpose detection/segmentation framework (see Figure 3), which has been successfully applied
to various field scenarios for object detection, motion detection, and instance segmentation
because of its flexibility for task-specific applications. The overall design of this framework can
be divided into two sequential stages; each including a subnetwork: (a) object detection, where
R-CNN (Ren et al. 2015) is applied to detect the instances of the semantic objects belonging to
certain classes in the image, and (b) pixel-wise segmentation, where the Fully Convolutional
Network (FCN) (Long et al. 2015) is employed for semantic segmentation with each pixel being
classified as part of a particle or the background.

SEGMENTATION KERNEL PERFORMANCE ANALYSIS

The proposed neural network, pre-trained COCO model, illustrated in Figure 3 was further
trained using the labeled aggregate image database. To validate and visualize the performance of
the segmentation kernel, 56 labeled images were randomly selected to serve as the validation set.
The validation set typically is a benchmark for assessing the performance of the trained models,
since the images in the validation set have never been used during the training procedure. Model
performance on the validation set indicates the accuracy, generality and robustness of the model
when dealing with unseen images. The trained kernel takes images from the validation set,
processes them and outputs the segmentation results with each particle marked by a colored
mask and a class category. Figure 4 shows the segmentation results of the various types of
sample images from three different sources.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 429
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Model architecture of the Mask R-CNN framework composed of: (a) Region-
based Convolutional Neural Network (R-CNN), (b) Fully Convolutional Network (FCN).

As illustrated in Figure 4, the segmentation kernel in general has gained the knowledge to
accomplish the image segmentation task and achieve relatively robust performance on aggregate
images under various conditions. Most of the aggregate particles with clear boundaries are
detected and extracted successfully from the image. To quantitatively evaluate the overall
performance of the segmentation kernel, two commonly used statistical indices, completeness
and precision, are introduced. Completeness refers to the ratio between the number of segmented
particles and the number of ground-truth labeled particles. Precision, on the other hand, is
defined by calculating the percentage of overlap between the segmented particle region and the
corresponding ground-truth mask. Within the context of Mask R-CNN model framework, the
completeness metric is treated as the indicator for the object detection performance while the
precision is considered to assess the performance of the pixel-wise segmentation task.
The completeness and precision results of 25 randomly selected images from the validation
set are summarized in Table 2. The average completeness and precision values are 79.7% and
86.7%, respectively. The precision of the segmentation kernel is very accurate, which indicates
the high quality of the mask prediction. The completeness of the kernel is relatively low, more
than 20% of the ground-truth particles are skipped by the model. The imbalance between the
precision and completeness could be eliminated by manually adjusting the confidence level
during inference time. Additionally, enriching the database or improving the architecture design
of the segmentation kernel is expected to improve the performance of the model. Finally, the
standard deviation values for completeness and precision are 7.8% and 4.5%, respectively as
seen in Table 2. These small numbers demonstrate the robustness of the model performance
under different conditions.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 430
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

(c) (d)

(e) (f)

Figure 4. Raw images collected from (a) an aggregate producer, (c) a railroad site and (e) a
laboratory setup, and the corresponding segmented images (b, d, and f).

Table 2. Completeness and Precision Results of Randomly Selected Validation Set Images

Completeness (%) Precision (%)


Average 79.7 86.7
Standard Deviation 7.8 4.5

AGGREGATE IMAGE ANALYSES AND DISCUSSION

After successful segmentation, each region represents a different aggregate particle within
the image, which is then fed into the post-processing module to calculate the outputs. For large-
sized riprap aggregate materials, the equivalent sizes and flat and elongated (F&E) ratios are
calculated for the segmented particles and are presented as a cumulative distribution. The

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 431

equivalent size of a particle follows the definition of the equivalent spherical diameter (ESD),
which is widely used to characterize the size of an irregularly shaped object. To calculate the
ESD, the measured area of the object is calculated as A. The ESD is then defined as follows:

𝐴
𝐸𝑆𝐷 = 2 × √
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

To calculate the F&E ratio, the maximum Feret dimension, 𝐿𝑚𝑎𝑥 is first determined by
searching for the longest intercept with the particle region in all possible directions. Next, by
searching the intercepts along the orthogonal directions against the 𝐿𝑚𝑎𝑥 , the minimum Feret
dimension, 𝐿𝑚𝑖𝑛 , is obtained. The F&E ratio is defined as follows:

𝐿𝑚𝑎𝑥
𝐹&𝐸 =
𝐿𝑚𝑖𝑛

As illustrated in Figure 5, both analyses show reasonable statistical distributions for the
morphological properties in an aggregate image and achieve good agreement with the ground-
truth labeling.

(a) (b)

Figure 5. Cumulative distribution curves for (a) equivalent particle size and (b) flat and
elongated ratio.

For medium-sized ballast aggregate particles, the PDS value is calculated based on the area
of the segmented particle and the total area of the input image according to the following steps:
(1) the area of each segmented particle and the total area of the image are first computed, which
are denoted as 𝑆𝑖 and 𝑆𝑖𝑚𝑎𝑔𝑒 , respectively; (2) the shortest dimension, 𝐷𝑖 , of each segmented
particle is extracted by searching for the shortest intercept with the particle region in all possible
directions; (3) the area of the particles whose shortest dimensions rage from 3/8 in. to 3 in. are
summed up to obtain the particle area, and (4) the PDS value is finally calculated by subtracting
from 100% the division of total particle area by the image canvas area. The PDS is defined as
follows:

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 432

∑ 3/8 𝑖𝑛.≤ 𝐷𝑖 ≤3 𝑖𝑛. 𝑆𝑖


𝑃𝐷𝑆 (%) = 100% −
𝑆𝑖𝑚𝑎𝑔𝑒

The PDS value for 26 field ballast aggregate images with known FI were calculated through
the proposed analysis pipeline. A statistical regression analysis (see Figure 6) was conducted to
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

establish a practical ballast degradation model relating FI to PDS values. The developed model
provided a linear relationship between the FI and PDS values with a coefficient of determination
(R2) equal to 0.72. The slope of the linear relationship is close to 1, which implies the robustness
of the model due to its non-sensitivity towards the change in both FI and PDS.

Figure 6. Ballast degradation model based on regression analyses for the 26 field ballast
image datasets with ground-truth FI values.

Note that the post-processing analysis presented in this paper is an example analysis with
simplified analytical components. Users should be attentive to the following aspects during a
formal and comprehensive analysis step. First, it is highly recommended for users to collect
images from a perpendicular direction against the target stockpile slope. The images in the
training dataset have no restrictions on viewing angle since they are used for the development
segmentation kernel. But the images for the post-processing stage and generating morphological
results should be normal-facing to minimize the perspective distortion effect on aggregate in the
images. Second, the segemented particle regions enable the user to conduct advanced analysis
for a more comprehensive characterization of particle shape in relation to the form, angularity
and texture. Finally, as a limitation, since only the surface of the stockpile is visible to users, it
should be noted that the analysis results only represent the aggregate statistics for the surface
particles in a stockpile.

CONCLUSIONS

A field aggregate evaluation system was designed for collecting and processing aggregate
images on site. A workflow involving preparation-training-analysis pipeline was established for
image segmentation and post-processing analysis. The state-of-the-art image segmentation
framework, named Mask R-CNN, was implemented with its segmentation kernel to accomplish

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 433

the challenging aggregate image segmentation task. By establishing a task-specific aggregate


image database, the trained segmentation model achieved good performance in terms of
segmenting out individual particles in an automated manner. Based on the findings of this study,
the proposed deep learning-based aggregate field evaluation procedure can serve as an efficient
and reliable tool for aggregate shape and size evaluation in the field.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ACKNOWLEDGEMENTS

This study was made possible through financial support provided by the Federal Railroad
Administration (FRA) and Illinois Department of Transportation (IDOT). The guidance and help
provided by Hugh Thompson, Ted Sussmann and Andrew J. Stolba are greatly appreciated.
Especially, the author would like to acknowledge railroad partners BNSF Railway, Canadian
National (CN) Railway and Amtrak, , Loram Maintenance of Way, Inc., and aggregate producers
RiverStone Group, Praire Materials, and Vulcan Materials companies for their help and support.

REFERENCES

Fowler, D. W., and Quiroga, P. N. (2003). The effects of aggregate characteristics on the
performance of Portland cement concrete. International Centre for Aggregate Research,
ICAR, 104.
Wnek, M. A., Tutumluer, E., Moaveni, M., and Gehringer, E. (2013). Investigation of aggregate
properties influencing railroad ballast performance. Transportation research record, 2374(1),
180-189.
Tutumluer, E., and Pan, T. (2008). Aggregate morphology affecting strength and permanent
deformation behavior of unbound aggregate materials. Journal of materials in civil
engineering, 20(9), 617-627.
Lagasse, P. F. (2006). Riprap design criteria, recommended specifications, and quality control
(Vol. 568). Transportation Research Board.
Barrett, P. J. (1980). The shape of rock particles, a critical review. Sedimentology, 27(3), 291-303.
Stark, T. D., and Wilk, S. T. (2018). Sampling, reconstituting, and gradation testing of railroad
ballast. Railroad Ballast Testing and Properties, ASTM STP1605, 125-133.
Huang, H., Luo, J., Tutumluer, E., Hart, J. M., and Stolba, A. J. (2020). Automated segmentation
and morphological analyses of stockpile aggregate images using deep convolutional neural
networks. Transportation Research Record, 2674(10), 285-298.
Huang, H., Luo, J., Tutumluer, E., Hart, J. M., and Qamhia, I. (2020). Size and shape
determination of riprap and large-sized aggregates using field imaging. Illinois Center for
Transportation/Illinois Department of Transportation.
Masad, E., Al-Rousan, T., Bathina, M., McGahan, J., and Spiegelman, C. (2007). Analysis of
aggregate shape characteristics and its relationship to hot mix asphalt performance. Road
Materials and Pavement Design, 8(2), 317-350.
Gates, L., Masad, E., Pyle, R., and Bushee, D. (2011). Aggregate imaging measurement system 2
(AIMS2) (No. FHWA-HIF-11-030).
Tutumluer, E., Rao, C., and Stefanski, J. A. (2000). Video image analysis of aggregates.
Moaveni, M., Wang, S., Hart, J. M., Tutumluer, E., and Ahuja, N. (2013). Evaluation of
aggregate size and shape by means of segmentation techniques and aggregate image
processing algorithms. Transportation research record, 2335(1), 50-59.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 434

He, K., Gkioxari, G., Dollár, P., and Girshick, R. (2017). Mask r-cnn. In Proceedings of the
IEEE international conference on computer vision (pp. 2961-2969).
Tutumluer, E., Ahuja, N., Hart, J. M., Moaveni, M., Huang, H., Zhao, Z., and Shah, S. (2017).
Field evaluation of ballast fouling conditions using machine vision (No. Safety- 27).
Selig, E. T., and Waters, J. M. (1994). Track geotechnology and substructure management.
Thomas Telford.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Dutta, A., Gupta, A., and Zissermann, A. (2016). VGG image annotator (VIA). URL:
http://www. robots. ox. ac. uk/vgg/software/via, 2.
Ren, S., He, K., Girshick, R., and Sun, J. (2015). Faster r-cnn: Towards real-time object
detection with region proposal networks. Advances in neural information processing systems,
28.
Long, J., Shelhamer, E., and Darrell, T. (2015). Fully convolutional networks for semantic s
egmentation. In Proceedings of the IEEE conference on computer vision and pattern
recognition (pp. 3431-3440).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 435

Computed Tomography of Sand Subjected to Heating: Analysis of Particle Displacements

Yize Pan1; Dawa Seo, Ph.D.2; Mark Rivers, Ph.D.3; Giuseppe Buscarnera, Ph.D.4;
and Alessandro F. Rotta Loria, Ph.D.5
1
Dept. of Civil and Environmental Engineering, Northwestern Univ., Evanston
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(corresponding author). Email: yizepan2018@u.northwestern.edu


2
Dept. of Civil and Environmental Engineering, Northwestern Univ., Evanston.
Email: dawaseo2021@u.northwestern.edu
3
Center for Advanced Radiation Sources, Univ. of Chicago, Argonne, IL.
Email: rivers@cars.uchicago.edu
4
Dept. of Civil and Environmental Engineering, Northwestern Univ., Evanston.
Email: g-buscarnera@northwestern.edu
5
Dept. of Civil and Environmental Engineering, Northwestern Univ., Evanston.
Email: af-rottaloria@northwestern.edu

ABSTRACT

Over the past decades, multiple studies have associated the deformation of coarse-grained
soils subjected to temperature variations with complex mechanisms involving particle
interactions at the microscale. Despite increasing advances in the understanding of how coarse-
grained soils subjected to temperature variations deform at the macroscale, knowledge at the
microscale remains limited, with arguably no experimental observations about such particle
interactions. This work presents an experimental investigation on particle interactions developing
in coarse-grained soils subjected to heating by means of X-ray computed microtomography.
Specifically, this work discusses a tomography experiment where particles of F-35 silica sand are
subjected to a temperature variation from 30C to 80C under zero applied vertical stress in a
small container. The displacements of individual particles are analyzed by image analyses on two
image scans before and after the heating. The results show that tomography experiments
combined with image analyses can provide sufficient accuracy at the microscale to capture
particle displacements induced by temperature variations. Heating induces upward particle
displacements whose magnitudes increase for particles located at shallower locations in the
assembly, suggesting a macroscopic volumetric expansion of sand upon heating that results from
the expansion and interactions of individual particles.

INTRODUCTION

Coarse-grained soils, like many other granular materials, are continuously subjected to
temperature variations due to natural or anthropogenic perturbations. Temperature variations can
have significant impacts on coarse-grained soils, with consequent detrimental influences on the
performance of engineering structures and energy geotechnologies (Laloui & Rotta Loria, 2019).
Therefore, understanding these impacts is paramount. In recent years, an increasing number of
experimental and numerical studies have analyzed the thermally induced deformation of coarse-
grained soils in the temperature range between 2°C and 200°C. Some experimental evidence
suggests that coarse-grained soils subjected to heating exhibit the thermal collapse phenomenon,
i.e., a macroscopic volumetric contraction of the granular assembly upon heating for even

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 436

minimal temperature variations (Agar, 1984; Agar et al., 1986; Kosar, 1983; Ng et al., 2016;
Sittidumrong et al., 2019; He et al., 2021). In contrast, other experimental studies show only
volumetric expansion upon heating (Ng et al., 2016; Liu et al., 2018, 2020; Pan et al., 2020,
2022), supporting complementary numerical evidence achieved via particle scale simulations
(Vargas & McCarthy, 2007; Dreissigacker et al., 2010; Sassine et al., 2018; Iliev et al., 2019;
Coulibaly et al., 2020; Zhao et al., 2020).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

One key limitation characterizing the current knowledge of the thermally induced
deformation of coarse-grained soils at the macroscale is the lack of experimental evidence about
the underlying mechanisms governing such phenomenon at the microscale. Previous studies have
indicated that the volumetric deformation of coarse-grained soils at the macroscale is triggered
by a modification of their structure at the microscale, resulting from the thermal deformation of
the individual particles that alters the contact force network and causes particle rearrangements.
However, such studies are purely numerical in nature and refer to granular assemblies composed
of spherical particles only, instead of considering actual coarse-grained soils (Vargas &
McCarthy, 2007; Dreissigacker et al., 2010; Sassine et al., 2018; Iliev et al., 2019; Coulibaly et
al., 2020; Zhao et al., 2020). Therefore, the full understanding of the actual thermally induced
deformation of coarse-grained soils remains limited, with arguably no experimental studies
available in this scope at the microscale.
Looking at such knowledge gap, this paper discussed some results of an X-ray computed
microtomography (CT) experimental campaign that has arguably been performed for the first
time to visualize and monitor the microstructural changes of coarse-grained soils subjected to
temperature variations. In this light of the unprecedented character of this endeavor, the simplest
possible experimental condition of dry sand subjected to heating under no applied vertical stress
in a container was considered. In the following, a succinct description of the beamline facility,
experimental setup, test material, image analyses, and results associated with this experimental
campaign is presented.

METHODOLOGY

In this section, the details of the methodology employed in this work are presented. The
beamline facility and experimental setup are introduced first. Then, the test material and
program, image analyses, and displacement calculations are presented.
Beamline Facility. Experiments were performed at the synchrotron microtomography
facility of Sector 13-BM-D GeoSoilEnviroCARS (GSECARS) at the Advanced Photon Source,
Argonne National Laboratory. With the pink beam tomography (Rivers, 2016), high-resolution
X-ray CT, images allowing for quantitative analyses at the scale of individual particle can be
collected. The spatial resolution of images is 3.13 μm per pixel. The temperature in the
tomography room is controlled constant at 23.51 °C.
The collection of CT images consists of a series of scanning movements, including (i)
rotation, (ii) vertical translation, and (iii) horizontal translation. The reason for each type of these
movements can be justified as follows. (i) The sample needs to be rotated 180 at acquisition
speed to collect the projections required for 3D reconstruction, then return to the original angle at
return speed. (ii) As the sample height exceeds the vertical field of view, one measurement will
require three datasets collected at three different heights. (iii) At each height, the sample needs to
be moved horizontally out of the view to collect the flat field for image correction. With the
consideration that these movements can potentially cause dynamic effects on the sample,

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 437

controlling their speeds and acceleration times appear to be critical for this experiment. Thus,
very slow speeds and long acceleration times as listed in Table 1 are used to ensure the
smoothest movements.
For each scan, 3600 projections are collected with a step angle of 0.05° per projection.
Depending on the rotation speed and number of projections, the exposure time is 0.022 s for each
projection. A mirror angle of 1.8 mrad and 0.25 mm Cu filter are selected to provide an optimal
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

contrast.

Table 1. Speeds and acceleration times during scanning movements

Rotation Horizontal translation Vertical translation


Acquisition Acceleration Return Max Acceleration Max Acceleration
speed [°/s] time [s] speed speed time [s] speed time [s]
[°/s] [mm/s] [mm/s]
2.2 2 3 1 5 1 5

Experimental setup. Temperature-controlled experiments are performed under zero applied


vertical stress in a container with no restraint at the top. Fig. 1 shows the experimental setup. The
sample container is made of fused quartz due to its low x-ray attenuation and very low thermal
expansion (linear thermal expansion coefficient of 5.5 × 10−7 /°C). The container has one flat
closed end with an inside diameter of 5 mm, an outside diameter of 7 mm, and an inside height
of 10 mm. No vertical mechanical load is applied to the sample. One single set screw is used to
hold the container at the base. A high-temperature polymer cell is built by 3D printer to provide
thermal insulation. A stream of nitrogen gas is blown from the top ceramic tube which contains
an electric furnace heater to heat the gas stream. The power output to the heater is controlled by a
PID temperature controller and the feedback temperature is read from a thermocouple attaching
the sample cylinder at the base.

Figure 1. Experimental setup, (a) global view, (b) detailed schematic cross section of the cell
with temperature control

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 438

Test Material and Program. The tested material is F-35 silica sand (Ottawa, IL) purchased
from U.S. Silica. It is characterized by a rounded shape and grain composition of over 99%
quartz (SiO2). The minimum and maximum void ratios of the sand are 𝑒𝑚𝑖𝑛 = 0.510 and 𝑒𝑚𝑎𝑥 =
0.764, respectively. The sand is artificially sieved to its mean size, between No. 30 (600 μm) and
No. 40 (425 μm) sieves. A loose sand sample is prepared with the air-pluviation method (Vaid &
Negussey, 1988), i.e., by raining uniformly the material into the container. The container is filled
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

below the top, at a height of around 9 mm, to avoid potential overflow during thermal expansion.
Two sets of images are collected at the temperatures of 30°C and 80°C. Fig. 2 shows the
setpoint temperature and feedback temperature reading. A heating rate of 0.1 °C/s for the
setpoint temperature is applied to the PID temperature controller. Since heat is generated by the
heater and circulated by the hot gas during heating, the feedback temperature is nearly
synchronized with the setpoint. Heating takes approximately 10 minutes for a temperature
increase of 50°C. At 80°C, the temperature stability of feedback temperature is  2°C due to
overshoot and fluctuation during scanning movements. At each temperature step, one set of
images is collected, which takes approximately 10 mins.

Figure 2. Setpoint and feedback (logged after heating from 30°C starts) temperatures

Image analyses. To analyze particle displacements, image analyses are performed on the
reconstructed CT images obtained from tomography experiments. Image analyses can be broken
down into two phases, termed as image processing and particle tracking, which are illustrated in
Fig. 3 and detailed below:
• Image processing: The raw tomography data is pre-processed and reconstructed through
the IDL tomoRecon software (Rivers, 2012). Then the reconstructed data is read and
processed through the Avizo software (FEI Company, 2019), which allows to reduce
image noise (filtering), segment the filtered image into solid and void phases
(binarization), separate individual particles (segmentation), label each particle in three
dimensions (labelling). After processing, the information of geometry (i.e., volume,
length, width, etc.) and position (i.e., center position in X, Y, Z directions as defined in
Fig. 4, etc.) for individual particles in each image set with a unique label index is output
from the Avizo software.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 439

• Particle tracking: To track individual particles between the two image sets, a tracking
algorithm (Seo et al., 2021) is used for analyzing the output from Avizo. Particle tracking
is achieved by matching the volumes and positions of particles. After matching the label
indices of particles in the two scans, displacements of individual particles can be
analyzed based on their center positions.
Further details of image processing via Avizo and particle tracking by the tracking algorithm
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

are discussed by Seo et al., (2021). Their approach to track particle breakage is not used in the
present work because such phenomenon does not characterize the tested material subjected to its
self-weight only.

Fig. 3. Illustration of image processing and particle tracking

Displacement calculation. When calculating particle displacements, it is important to set up


the coordinate system to ensure a consistent reference origin and axis directions. This is due to
the unavoidable translation and tilt of the sample cylinder during scanning movements and
thermal expansion. (i) Origin centering and (ii) tilt correction of coordinate system are performed
as shown in Figure 4 to correct for the potential translation and tilt of the cylinder, respectively.
(i) The origin of coordinate system in each image set is calculated by averaging the positions of
particles at the bottom layer (one layer), assuming negligible relative movements between those
particles and the cylinder base. (ii) For every two image sets, the coordinate system (X, Y, Z) of
the current image set is corrected for cylinder tilt angles 𝜃𝑋𝑍 and 𝜃𝑌𝑍 in both the XZ and YZ
planes compared to the initial image set. Considering the XZ plane for example, the tilt angle
𝜃𝑋𝑍 is obtained from the slope angle when plotting Z displacement (uncorrected) 𝑍 – 𝑍0 versus
initial position 𝑋0 of bottom layer particles. Thus, the corrected positions in X and Z directions in
the current image set are given by:

𝑋 ′ = 𝑋𝑐𝑜𝑠𝜃𝑋𝑍 + 𝑍𝑠𝑖𝑛𝜃𝑋𝑍 (1)

𝑍 ′ = −𝑋𝑠𝑖𝑛𝜃𝑋𝑍 + 𝑍𝑐𝑜𝑠𝜃𝑋𝑍 (2)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 440

The correction in the YZ plane is applied following the same principle above. The results of
cylinder tilt will be discussed in the next section. The corrected displacements of particles in the
Z direction can be expressed as:

∆𝑍 = 𝑍 ′ − 𝑍0 (3)
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Centering and tilt correction (consider XZ plane for example) of coordinate system

RESULTS AND DISCUSSIONS

In this section, the results of the quantitative analyses on particle displacements for sand
subjected to heating are presented and discussed. In this context, the information of cylinder tilt,
which is corrected in the final results of particle displacements, is presented first. Then, the
corrected particle displacements induced by a temperature increase of 50°C is discussed.
Cylinder tilt. Fig. 5 shows the plots of particle displacements in Z direction versus their
initial positions in horizontal directions, 𝑋0 and 𝑌0 , for the bottom layer particles. The slopes in
the plots before correction suggest a tilt of cylinder between the two image sets collected at 30
°C and 80 °C. The approximate tilt angles are 𝜃𝑋𝑍 = −0.304° in the XZ plane, and 𝜃𝑌𝑍 =
0.189° in the YZ plane. The slopes become minimal (one order smaller) after tilt corrections.
The main cause for this evidence might be the thermal expansion of the system (e.g., the screw
and cylinder, etc.) upon heating. Such cylinder tilt can result in a significant error up to 0.015
mm in the results of particle displacements in Z direction, thus tilt corrections are essential for
the analyses of particle displacements when subjected to temperature variations.
Particle displacements. Fig. 6 shows the corrected displacements along the Z direction that
characterize the particles of monodisperse F-35 sand when subjected to heating. The plot adopts
the sign convention that considers vertical settlements (downward displacements) as negative
according to the coordinate system.
The results show that 95% of the particles undergo upward displacements upon heating,
suggesting a macroscopic volumetric expansion of the sample. An overall increase in the
magnitudes of displacements suggests that the particles located at shallower locations in the
sample experience larger upward displacements. This evidence is attributed to the lower restraint
characterizing the particles at these locations compared to deeper locations. Two discontinuities
in distribution of vertical displacements along the sample height are observed, with 𝑍0 = 3 and 6

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 441

mm. This evidence is associated to potential errors at the image stitches as three separate images
at different heights are combined, especially when considering the possible thermally induced
inaccuracies of position control, as well as the temporal fluctuations and spatial variations of
temperature field. Another possible cause for the discontinuities may be the disturbances by
scanning movements, which cannot be avoided even when very slow speeds and long
acceleration times are adopted as in the present study, especially since no vertical mechanical
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

load is applied to the sample.

Fig. 5. Cylinder tilt indicated by the bottom layer particles

Fig. 6. Particle displacements in Z direction (after correction) when subjected to heating of


50°C

The results obtained in this study at the microscale corroborate those previously reported by
Pan et al., (2020) at the macroscale, where loose sand subjected to minimal vertical stress

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 442

exhibited volumetric expansion upon heating. In addition, the current work highlights for the
first time the individual particle displacements that characterize coarse-grained soils subjected to
heating, suggesting a significant role of local restraint conditions and perturbations on the overall
displacement distributions of particles. The methodology introduced in this work, which
combines tomography experiments and image analyses, specifically appears capable of
providing sufficient accuracy to quantitatively capture the particle displacements of coarse-
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

grained soils at the microscale, with rich potential for future studies on the mechanics of coarse-
grained soils under non-isothermal conditions and the consideration of even more complex
thermo-mechanical loading paths.

CONCLUSION

In this paper, the displacements of individual particles constituting monodisperse F-35


silica sand subjected to heating under zero applied vertical stress in a container are
investigated by tomography experiments and image analyses. The main conclusions of this
work are as follows:
• Heating causes a non-uniform distribution of vertical displacements from the bottom to
the top of the tested material, with particles that move more at locations involving lower
localized restraint conditions (i.e., at shallower depths in the sample).
• The distribution of vertical displacement can suffer from perturbations and errors during
testing, which can be responsible for localized inversions in the direction of
displacements.
• The dominant upward displacements of particles upon heating suggest that the tested
coarse-grained soil macroscopically expands upon heating, with no thermal collapse.
• CT tomography experiments and image analyses can provide sufficient accuracy to
perform quantitative analyses on the particle displacements of sand when subjected to
temperature variations at the microscale. Such methodology can be further expanded in
the future to study the microstructural changes of coarse-grained soils subjected to
temperature variations under more complex conditions, e.g., non-zero vertical stress.
• Error control is essential for experiments such as those presented in this work. A
consistent coordinate reference system is critical for analyzing particle displacements.
Imperfections in position and temperature controls, and scanning movements generating
dynamic disturbances, might result in errors.

ACKNOWLEDGEMENT

The first and last authors of this work are grateful to the support provided by the United
States Army Research Office to carry out this study (project grant W911NF2110059). This work
was performed at GeoSoilEnviroCARS (The University of Chicago, Sector 13), Advanced
Photon Source (APS), Argonne National Laboratory. GeoSoilEnviroCARS is supported by the
National Science Foundation – Earth Sciences (EAR – 1634415). This research used resources of
the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User
Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract
No. DE-AC02-06CH11357.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 443

REFERENCES

Agar, J. G. (1984). Geotechnical behaviour of oil sands at elevated temperatures and pressures
[Ph.D. Thesis]. Univerity of Alberta.
Agar, J. G., Morgenstern, N. R., and Scott, J. D. (1986). Thermal-expansion and pore pressure
generation in oil sands. Canadian Geotechnical Journal, 23(3), 327–333.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Coulibaly, J. B., Shah, M., and Rotta Loria, A. F. (2020). Thermal cycling effects on the
structure and physical properties of granular materials. Granular Matter, 22(4), 80.
Dreissigacker, V., Müller-Steinhagen, H., and Zunft, S. (2010). Thermo-mechanical analysis of
packed beds for large-scale storage of high temperature heat. Heat and Mass Transfer,
46(10), 1199–1207.
FEI Company. (2019). Avizo 3D image analysis software.
He, S.-H., Shan, H.-F., Xia, T.-D., Liu, Z.-J., Ding, Z., and Xia, F. (2021). The effect of
temperature on the drained shear behavior of calcareous sand. Acta Geotechnica, 16(2), 613–
633.
Iliev, P. S., Giacomazzi, E., Wittel, F. K., Mendoza, M., Haselbacher, A., and Herrmann, H. J.
(2019). Behavior of confined granular beds under cyclic thermal loading. Granular Matter,
21(3), 59.
Kosar, K. M. (1983). The effect of heated foundations on oil sand [Master’s Thesis]. University
of Alberta.
Laloui, L., and Rotta Loria, A. F. (2019). Analysis and design of energy geostructures:
Theoretical essentials and practical application. Elsevier Academic Press.
Liu, H., Liu, H., Xiao, Y., and McCartney, J. S. (2018). Influence of temperature on the volume
change behavior of saturated sand. Geotechnical Testing Journal, 41(4), 747–758.
Liu, H., McCartney, J. S., and Xiao, Y. (2020). Thermal volume changes of saturated sand
during loading-unloading-heating phase. E3S Web of Conferences, 205, 08002.
Ng, C. W. W., Wang, S. H., and Zhou, C. (2016). Volume change behaviour of saturated sand
under thermal cycles. Géotechnique Letters, 6(2), 124–131.
Pan, Y., Coulibaly, J. B., and Rotta Loria, A. F. (2020). Thermally induced deformation of
coarse-grained soils under nearly zero vertical stress. Géotechnique Letters, 10(4), 486–491.
Pan, Y., Coulibaly, J. B., and Rotta Loria, A. F. (2022). An experimental investigation
challenging the thermal collapse of sand. Géotechnique, 1–27.
Rivers, M. L. (2012). tomoRecon: High-speed tomography reconstruction on workstations using
multi-threading. Developments in X-Ray Tomography VIII, 8506, 85060U.
Rivers, M. L. (2016). High-speed tomography using pink beam at GeoSoilEnviroCARS. 9967.
Sassine, N., Donzé, F.-V., Harthong, B., and Bruch, A. (2018). Thermal stress numerical study in
granular packed bed storage tank. Granular Matter, 20(3), 1–15.
Seo, D., Sohn, C., Cil, M. B., and Buscarnera, G. (2021). Evolution of particle morphology and
mode of fracture during the oedometric compression of sand. Géotechnique, 71(10), 853–
865.
Sittidumrong, J., Jotisankasa, A., and Chantawarangul, K. (2019). Effect of thermal cycles on
volumetric behaviour of Bangkok sand. Geomechanics for Energy and the Environment,
100127.
Vaid, Y. P., and Negussey, D. (1988). Preparation of Reconstituted Sand Specimens. Advanced
Triaxial Testing of Soil and Rock.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 444

Vargas, W. L., and McCarthy, J. J. (2007). Thermal expansion effects and heat conduction in
granular materials. Physical Review E, 76(4), 041301.
Zhao, S., Zhao, J., and Lai, Y. (2020). Multiscale modeling of thermo-mechanical responses of
granular materials: A hierarchical continuum–discrete coupling approach. Computer Methods
in Applied Mechanics and Engineering, 367, 113100.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 445

Pre-Drilling Effects on Vibrations and Ground Deformations Caused by Impact Pile


Driving

Berk Turkel, S.M.ASCE1; Jorge E. Orozco-Herrera, S.M.ASCE2;


Luis G. Arboleda-Monsalve, Ph.D., M.ASCE3; Boo Hyun Nam, Ph.D., P.E.4; and Larry Jones5
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Graduate Research Assistant, Dept. of Civil, Environmental, and Construction Engineering,
Univ. of Central Florida, Orlando, FL. Email: TurkelBerk@knights.ucf.edu
2
Graduate Research Assistant, Dept. of Civil, Environmental, and Construction Engineering,
Univ. of Central Florida, Orlando, FL. Email: Jeorozcoh@knights.ucf.edu
3
Assistant Professor, Dept. of Civil, Environmental, and Construction Engineering, Univ. of
Central Florida, Orlando, FL. Email: Luis.Arboleda@ucf.edu
4
Associate Professor, Dept. of Civil, Environmental, and Construction Engineering, Univ. of
Central Florida, Orlando, FL. Email: BooHyun.Nam@ucf.edu
5
State Geotechnical Engineer, Florida Dept. of Transportation, Tallahassee, FL.
Email: Larry.Jones@dot.state.fl.us

ABSTRACT

Driven precast concrete piles are used as a compelling deep foundation alternative and can be
installed below the groundwater table without providing any casing and also unnecessary time
delays. Granular materials with relative densities ranging from loose to medium-dense conditions
constitute an ideal scenario to install pile foundations using impact driving methods since their
bearing capacity might increase due to soil compaction. However, those soil conditions can be
highly susceptible to ground deformations induced by pile driving operations mainly due to
particle rearrangement and excess pore water pressure buildup and dissipation. Pre-drilling can
play a significant role in pile driving-induced deformations and vibrations since the waves
emanating from the pile get attenuated prior to reaching the ground surface as the distance
between the pile tip and the ground surface increases. This paper presents the effects of pre-
drilling installation procedures of prestressed concrete piles on the ground response quantified in
terms of deformations and vibrations using a finite element model conducted in PLAXIS 2D.
The authors analyzed two configurations of pre-drilling depths selected from bridge construction
projects in Florida. The critical-state based hypoplasticity model for sands enhanced with the
intergranular strain concept was used as the constitutive soil model for the analyses, thus
allowing the program to track changes in the state of stresses and void ratios of the soil as the
pile is being driven in the ground in a continuous large-deformation approach. Two diesel
hammers and their corresponding forces applied to the top of the pile were used in the numerical
models to consider the effects of the selected rated energies in the final response. The results of
the numerical analyses indicated that pre-drilling effectively reduces both ground deformations
and vibrations quantified in terms of peak particle velocities due to the relative position of the
pile tip in relation to the distance to the ground surface.

INTRODUCTION

Ground vibrations induced by pile driving activities are usually associated with possible
structural damage to nearby infrastructure. It is common practice to quantify those vibrations in
terms of the Peak Particle Velocity (PPV). Regulatory agencies and construction codes normally

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 446

define PPV levels that cannot be exceeded in order to prevent structural damage. PPV reference
values from several sources (e.g., British Standards Institution, German Institute of Standards,
and U.S. Office of Surface Mining) are presented by Athanasopoulos and Pelekis (2000). In this
paper, a value of 12.7 mm/s was used as a reference PPV value as defined by the Florida
Department of Transportation (FDOT, 2021).
In addition to the inherent damages that ground vibrations can cause to nearby structures,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ground deformations (e.g., differential settlement, heave) associated with dynamic effects in the
soil can also be detrimental to structures as shown by Massarsch and Fellenius (2014). Orozco-
Herrera et al. (2022) showed that ground deformations due to soil densification can still occur
even if PPV levels are low.
Pre-drilling the piles represents a commonly used method to reduce pile driving-induced
vibrations. This can help reducing the ground surface vibrations by increasing the distance from
the pile tip to the ground surface. This effect has already been solely quantified for ground
vibrations in previous studies performed by Grizi et al. (2016) and Heung et al. (2007). FDOT
(2021) already incorporated pre-drilling effects into their guidelines by acknowledging its
beneficial effects on reducing ground vibrations due to pile driving.
This study aimed to determine the additional effect of pre-drilling on ground deformations by
numerically simulating a continuous pile driving on a finite element (FE) environment. The wave
equation analysis software GRLWEAP was also used to obtain the forces that selected impact
hammers apply at the top of the pile. Subsequently, those forces were applied into the numerical
analyses performed in the FE platform PLAXIS 2D. As shown by Turkel et al. (2021),
continuous pile driving (i.e., pile penetration modeled continuously up to a final target depth) is
preferable for the analysis of ground deformations because such an analysis framework is
capable of representing the accumulation of stresses and deformations during the pile driving
process.

NUMERICAL PILE DRIVING MODELS

Table 1 presents the two selected pre-drilling configurations considered in this paper. A pre-
drilling of approximately 9.7 m corresponding to a pre-drilling to pile length ratio of 35% for a
27.5 m long pile was selected based on a survey conducted by the authors in pile driving sites in
Central Florida. Model No. 1 considers a low pre-drilling ratio to analyze the detrimental effects
caused by pile tips being closer to the ground surface.

Table 1. Summary of analyses performed to elucidate pre-drilling effects on ground


response.

Pile Length Pre-Drilling Depth


Model Pre-Drilling Ratio*
(m) (m)
1 27.5 7.0 25%
2 27.5 9.7 35%
*Calculated as the ratio between the pre-drilling depth and the pile length.

GRLWEAP Pile Driving Model

The forcing functions that were applied to the FE models were first obtained from a wave
equation analysis performed in GRLWEAP. The selected impact hammers for the analyses were

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 447

the APE D70-52 and DELMAG D36-32 following the modeling approaches conducted by
Orozco-Herrera et al. (2022). These hammers are commonly used in the state of Florida as
reported by Heung et al. (2007). Table 2 presents the rated energy for each hammer. Notice that
the largest and lowest energies correspond to the APE D70-52 and DELMAG D36-32 hammers,
respectively. These hammers were also selected to allow the authors to analyze the pile driving
process in a wide range of rated energies and forcing functions that have a direct impact on the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

computed ground responses.

Table 2. Selected hammer types for the parametric studies including their rated energies.

Hammer Type Rated Energy (kJ)


APE D70-52 236.2
DELMAG D36-32 123.2

The analyses in GRLWEAP were performed by simulating a 610 mm-wide square precast
concrete pile at a final penetration depth of 27.5 m and a pre-drilling of 9.7 m (i.e., conditions of
model 2). The pile length was also defined as 27.5 m. The soil profile consisted of a surficial
31.2 m-thick medium-dense sandy layer underlain by a 53.0 m-thick very dense granular
stratum. Additionally, a 38.1 mm-thick hammer cushion and a 381 mm-thick pile cushion were
used. The material of the pile cushion was selected as “used” plywood to properly model the
thickness reduction due to multiple hammer blows, thus modeling larger transmitted energy.
Figure 1 presents the computed forcing functions applied by a single hammer blow for each
selected hammer. Notice that the largest computed peak force is applied by the APE D70-52
hammer.

9
APE D70-52
8 D36-32
7
Force (MN)

6
5
4
3
2
1
0

0.00 0.02 0.04 0.06 0.08


Time (s)

Figure 1. Computed forcing function by a single hammer blow to be applied at the top of
the pile in the finite element models.

Finite Element Modeling of Pile Driving

Figure 2a presents the model geometry and the idealized soil conditions. It has been
previously observed that the predominant soil conditions in Central Florida consist mainly of
loose to medium-dense soils (Heung et al., 2007; Bayraktar et al., 2013; and Turkel et al., 2021).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 448

The FE model consisted of a top 31.2 m-thick medium-dense sand with a relative void ratio (𝑟𝑒 )
of 55% underlain by a very dense sand with an 𝑟𝑒 of 90%. The relative void ratio is an analog
variable to the relative density of the material and it establishes the in-situ conditions with
respect to the critical state line as defined by Masin (2019). This bottom layer was introduced in
the model to represent a firm competent bearing stratum where pile driving operations are
normally completed. The total height of the mesh was 84.0 m while the total width was 94.0 m.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

The bottom boundary was set as fully fixed, and the right and left boundaries were defined as
normally fixed. The potential reflection of the emanating waves at the boundaries was avoided
by defining viscous boundary conditions at the bottom and right-hand side edges of the model.
The selected elements for the entire mesh were fifteen-node triangles. Figure 2b presents a
detailed view of the soil clusters of a refined mesh zone with a coarseness factor of 0.25 that was
introduced in the model to represent a near-zone of the pile driving operation of approximately
one pile length. The updated mesh option in PLAXIS 2D was activated in the analyses to
account for the large deformations that occur as a result of the continuous pile driving modeling
approach adopted in this paper. As mentioned by previous studies (e.g., Grizi et al., 2018 and
Turkel et al., 2021), the interface elements suggested by PLAXIS 2D can lead to numerical
instabilities when performing dynamic loading to the pile. This paper uses a critical state-based
constitutive soil model (i.e., Hypoplasticity for sands) that can lead to critical state conditions
close to the pile as it is being driven, thus allowing for large deformations with fully mobilized
strength conditions around the pile (i.e., similar to an interface element).

Figure 2. Pile driving model in PLAXIS 2D: (a) model geometry and (b) detailed view of
the refined zone and pre-drilling pile depth (model 1).

The forcing functions obtained from the wave equation analyses were applied to the FE
models to study the effects of selected hammers and pre-drilling depths on the computed ground
response (i.e., ground deformations and vibrations in terms of PPV). The pile driving was
modeled using a three-staged construction process. The first stage was defined to initialize the
field state of stresses in each soil layer. The second stage was defined to activate the pile cluster
at the specified pre-drilling depth (i.e., 7.0 m and 9.7 m). The third stage consisted of the
application of the hammer blows using computed distributed stresses at the top of the pile. Each
hammer blow was applied with a 1-second interval up to 1400 hammer blows.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 449

The behavior of the top medium-dense sandy layer was modeled with the hypoplasticity
model for sands developed by von Wolffersdorff (1996) and enhanced with the intergranular
strain concept by Niemunis and Herle (1997). This constitutive soil model was developed for
PLAXIS 2D by Gudehus et al. (2008). This model was selected since its critical state framework
allows the accurate computation and update of void ratios (e) and changes in volume as the pile
driving operation progresses. This is important for this case because ground deformations due to
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

pile driving can also be associated with soil densification (Orozco-Herrera, 2021).

Definition of the Soil Parameters

The relative void ratio of the top sandy layer was defined as 55%. Figure 3 presents
computed laboratory tests that were simulated to define the soil parameters. The definition of the
soil parameters was performed similarly to the work by Orozco-Herrera et al. (2022) via stress-
controlled undrained triaxial compression tests consolidated to 𝐾0 conditions (𝐶𝐾0 𝑈 − 𝑇𝑋𝐶)
which were also conducted using the soil testing module available in PLAXIS 2D (see Figures
3a and 3b). Both computed deviatoric stress (Δ𝑞) and excess pore water pressures (Δ𝑢) are
presented versus axial strains (𝜖𝑎 ). An initial cell pressure of 100 kPa and a 𝐾0 of 0.5 were
selected. A dilative shearing response was observed during the computed numerically-simulated
tests, which was the soil behavior expected for the range of applied confining pressures, also
confirmed by Hyodo et al. (1994). The parameters were defined to replicate expected shear
modulus degradation curves previously published by Hardin and Drnevich (1972) and Seed and
Idriss (1970) (see Figure 3c) for the selected relative void ratios. Notice that the shear modulus
degradation defined by Seed and Idriss (1970) involved both cyclic and monotonic tests.
Examples of the monotonic triaxial (𝑀𝑜𝑛. 𝑇𝑋) tests used to define these curves are presented in
Figures 3c for relative densities (𝐷𝑟 ) of 40% and 70%. The computed shear modulus degradation
curve of the soil was computed by following the methodology proposed by Arboleda-Monsalve
et al. (2017). The equation developed by Bauer (1996), which is based on granular hardness (ℎ𝑠 )
and an exponent for the grain skeleton (𝑛), was used to calculate the void ratio versus applied
pressure relationships since the void ratio at the specified mean confining pressure (𝑝𝑠 ) is
required to plot the reference degradation curves.

400 100
re=55%, e0=0.81 (a) (c)
Dq (kPa)

80
G (MPa)

300 Dr=75%
Dr=40%
200 60
p =133 kPa
s
40
100 ea (%)
0 1 2 3 4 5
20
0
-50 (b) 0
0.0001
1E-4 0.001
1E-3 0.01 0.1 1 10
-40 re=55%, e0=0.81 gs (%)
Du (kPa)

-30 PLAXIS: re=55%, e0=0.81


-20 Hardin and Drnevich (1972)
-10 Seed and Idriss (1970)
Seed and Idriss (1970) (Weissman and Hart, 1961)- Mon. TX
0 Seed and Idriss (1970) (Donovan, 1968)- Mon. TX
10 0 1 2 3 4 5 Seed and Idriss (1970) (Donovan, 1969)- Mon. TX
ea (%)

Figure 3. Computed results from simulated triaxial tests (𝑪𝑲𝟎 𝑼 − 𝑻𝑿𝑪): (a) 𝚫𝒒 versus 𝝐𝒂
(b) 𝚫𝒖 versus 𝝐𝒂 and (c) secant shear stiffness degradation curves for a relative void ratio
of 55%.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 450

The selected set of parameters used in the previous numerically simulated tests is shown in
Table 3. The 𝑒0 value of 0.81 was defined as corresponding to a relative void ratio of 55% and
reference values for the maximum void ratio (𝑒𝑚𝑎𝑥 𝑜𝑟 𝑒𝑑0 in the model) and minimum void ratio
(𝑒𝑚𝑖𝑛 𝑜𝑟 𝑒𝑐0 in the model) were defined based on the values reported for similar soil conditions
by Lade et al. (1998) and Zapata-Medina et al. (2019).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Table 3. Soil properties used for the Hypoplasticity sand model in PLAXIS 2D.

No. Parameter Description Value Unit


1 𝜙c Critical state friction angle 31 °
2 pt Shift of the mean stress due to cohesion 0 kPa
3 hs Granular hardness 1200 MPa
4 n Exponent for pressure sensitive of a grain skeleton 0.37 -
5 ed0 Minimum void ratio at zero pressure (ps = 0) 0.58 -
6 ec0 Critical void ratio at zero pressure (ps = 0) 1.096 -
7 ei0 Maximum void ratio at zero pressure (ps = 0) 1.315 -
Exponent for transition between peak and critical
8 α 0.05 -
stresses
Exponent for stiffness dependency on pressure and
9 β 1.4 -
density
10 𝑚𝑅 Stiffness increase for 180° strain reversal 5 -
11 𝑚𝑇 Stiffness increase for 90° strain reversal 2 -
12 Rmax Size of elastic range 5.00x10-5 -
13 βr Material constant representing stiffness degradation 0.1 -
Material constant for evolution of intergranular
14 χ 1.0 -
strains
15 𝑒0 Initial void ratio at zero pressure (𝑝𝑠 =0) 0.81

RESULTS OF THE NUMERICAL MODEL

Figure 4 presents the pile tip depth as the hammer blows are applied for the selected pre-
drilling depths of 7.0 m and 9.7 m (i.e., pre-drilling-to-pile length ratios of 25% and 35%,
respectively). Each curve starts from the corresponding pre-drilling depth. The analyses were
finalized when one of the following conditions occurred: the bearing stratum was reached, 1400
hammer blows were applied, or a lack of convergence or excessive computational time occurred
during the analyses. It can be observed that a shallower pre-drilling depth required more hammer
blows to reach the same depth than the deeper pre-drilling for both hammer types (see Figures 4a
and 4b, respectively). The penetration rate, quantified in terms of the slope of the curves, tends to
flatten more rapidly for the shallow pre-drilling depth. This can be attributed to the densification
of the surrounding soils when pile driving starts at a shallower depth (i.e., more hammer blows
cause vibration-induced soil densification in terms of a reduction of void ratio in critical state
space). In general, the computed results indicate that the type of hammer and its corresponding
input energy have a large influence on the ground response in terms of the studied evolution of
the pile penetration process.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 451

Hammer Blows Hammer Blows


0 200 400 600 800 1000 1200 1400 0 200 400 600 800 1000 1200 1400
0
(a) (b)

Pile Tip Depth (m)


10
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

15 Pre-drilling ratio 35%


Pre-drilling ratio 25%

20

25

30

Figure 4. Effect of pre-drilling on the computed vertical pile penetration through sandy soil
with a relative void ratio of 55% by using hammer types: (a) APE D70-52 and (b)
DELMAG D36-32.

Figure 5 presents the computed PPV attenuation curves at the ground surface for each type of
hammer and pre-drilling depths. The PPV attenuation curves computed by using an APE D70-
52, and DELMAG D36-32 hammers are presented in Figures 5a and 5b, respectively. Note that
the importance of pre-drilling is reflected in the fact that PPVs are smaller in the case of the large
pre-drilling depth (i.e., a pre-drilling ratio of 35%). The trend is noticeable when driving the piles
using the APE D70-52 hammer (see Figure 5a). The effect of the pre-drilling ratio is not as
significant for those scaled distances corresponding to PPV values lower than the previously
mentioned reference value of 12.7 mm/s. The trend computed for the selected hammers showed
that the larger the pre-drilling ratio, the lower the pile driving-induced ground vibrations.
However, the decrease in the PPV is not as significant as expected.

(a) Pre-drilling ratio 25% (b)


1000
Pre-drilling ratio 35%
Regression line-25% PPV=5.89(D/√E)-1.52
R2=0.97
Regression line-35%

PPV=3.55(D/√E) -1.86
R2=0.95
100
PPV (mm/s)

PPV=4.27(D/√E) -1.68
PPV=6.34(D/√E)-1.50
R2=0.94
R2=0.97
10

1
0.01 0.1 1 10 0.01 0.1 1 10
Scaled Distance (m/√kJ) Scaled Distance (m/√kJ)

Figure 5. Effect of pre-drilling depth on the computed PPV attenuation curves during pile
driving through a sandy soil with a relative density of 55% and for the hammer types: (a)
APE D70-52 and (b) DELMAG D36-32.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 452

Figure 6 presents the maximum computed ground surface deformations during the pile
driving for the selected hammers and pre-drilling ratios. The ground deformations were
computed up to a pile tip depth of 18.0 m regardless of the pre-drilling ratio or hammer used
since all the analyses were able to reach such depth. This was preferable to avoid discrepancies
associated with the application of more hammer blows for different hammers and pre-drilling
configurations. The ground deformation values computed by using APE D70-52 and DELMAG
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

D36-32 hammers are presented in Figures 6a and 6b, respectively. Notice that the ground
deformations decreased for the deeper pre-drilling ratio, regardless of the hammer used. This
trend is not very clear for those points close to the pile due to large soil disturbances in the
proximity to the pile. For points further away from the pile, the ground deformations are largely
affected by the pre-drilling ratio. This shows the benefit of providing, when possible, large pre-
drilling ratios to prevent ground deformations associated with pile driving. This can be attributed
to the larger length of the path that the emanating waves from the pile (particularly the spherical
waves resulting from the pile tip) have to travel until they become surface waves.

Distance from the pile (m) Distance from the pile (m)
0 5 10 15 20 0 5 10 15 20
60
(a) (b)
50
40
Ground deformation (mm)

30
20
Settlement
10
0
-10
Heave
-20
-30
-40
-50
Pre-drilling ratio 25% Pre-drilling ratio 35%

Figure 6. Effect of pre-drilling depth on the maximum computed ground surface settlement
and heave during pile driving through sandy soil with a relative void ratio of 55% and for
the hammer types: (a) APE D70-52 and (b) DELMAG D36-32.

SUMMARY AND CONCLUSIONS

This study presents the results of numerical simulations conducted in PLAXIS 2D to analyze
the effects of pre-drilling ratios on pile driving-induced ground responses. The analyses were
performed using the hypoplasticity model for sands to simulate soil densification since its critical
state framework allows the update of void ratios (e) and changes in volume as the pile driving
operation progresses. Two pre-drilling ratios corresponding to 25% and 35% were analyzed. The
following conclusions are drawn from this paper:
1. The selection of a pre-drilling ratio is an important factor in potentially attenuating
detrimental vibrations caused by pile driving operations. This was quantified in terms of
the computed PPVs and ground deformations.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 453

2. Pile driving penetration rates are highly dependent on the type of hammer and pre-
drilling ratio. For deeper pre-drillings, the effort required to drive the pile might decrease
and vice-versa. This can be attributed to changes in volume and soil densification due to a
large number of hammer blows.
3. The results showed as expected that the larger selected pre-drilling ratio caused smaller
ground deformations. This can be attributed to the previous conclusion showing that soil
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

densifies more for shallower pre-drillings, thus increasing the ground deformations.
4. This trend was not that clear in the case of PPVs. Pre-drilling affects PPVs due to the fact
that the path that the emanating waves from the pile have to travel up to the ground
surface is increased. Further numerical analyses are required to elucidate the effects of
pre-drilling on PPVs on the ground surface.

ACKNOWLEDGEMENTS

Financial support was provided by the Florida Department of Transportation, Project No.
BDV24 TWO 977-33. The opinions, findings, and conclusions expressed in this publication are
those of the authors and not necessarily those of the Florida Department of Transportation or the
U.S. Department of Transportation.

REFERENCES

Arboleda-Monsalve, L. G., Teng, F., Kim, T., and Finno, R. J. 2017. Numerical Simulation of
Triaxial Stress Probes and Recent Stress-History Effects of Compressible Chicago Glacial
Clays. J. Geotech. Geoenviron. Eng. 143, 04017029.
Athanasopoulos, G. A., and Pelekis, P. C. 2000. Ground vibrations from sheet pile driving in
urban environment: measurements, analysis and effects on buildings and occupants. Soil
Dynamics and Earthquake Engineering 19, 371–387.
Bauer, E. 1996. Calibration of a Comprehensive Hypoplastic Model for Granular Materials. Soils
and Foundations 36, 13–26.
Bayraktar, M. E., Kang, Y., Svinkin, M., and Arif, F. 2013. Evaluation of Vibration Limits and
Mitigation Techniques for Urban Construction. Florida Department of Transportation,
Tallahassee, FL.
FDOT (Florida Department of Transportation). 2021. Section 455: Structures Foundations, in:
Standard Specifications for Road and Bridge Construction. Florida Department of
Transportation (FDOT), Tallahassee, FL, pp. 552–619.
Grizi, A., Athanasopoulos-Zekkos, A., and Woods, R. D. 2018. H-Pile Driving Induced
Vibrations: Reduced-Scale Laboratory Testing and Numerical Analysis, in: IFCEE 2018.
Presented at the IFCEE 2018, American Society of Civil Engineers, Orlando, Florida, pp.
165–175.
Grizi, A., Athanasopoulos-Zekkos, A., and Woods, R. D. 2016. Ground Vibration Measurements
near Impact Pile Driving. J. Geotech. Geoenviron. Eng. 142, 04016035.
Gudehus, G., et al. 2008. The soilmodels.info project. Int. J. Numer. Anal. Meth. Geomech. 32,
1571–1572.
Hardin, B. O., and Drnevich, V. P. 1972. Shear Modulus and Damping in Soils: Design
Equations and Curves. J. Soil Mech. and Found. Div. 98, 667–692.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 454

Heung, W., Morgan, K., Yoon, Y. H., Gobin, R., and Gollamudi, S. 2007. Vibration due to
Driving Concrete Piles Using Open-Ended Diesel Hammer in Central and South Florida, in:
7th FMGM 2007. Presented at the Seventh International Symposium on Field Measurements
in Geomechanics, American Society of Civil Engineers, Boston, MA, pp. 1–12.
Hyodo, M., Tanimizu, H., Yasufuku, N., and Murata, H. 1994. Undrained cyclic and monotonic
triaxial behaviour of saturated loose sand. Soils and Foundations 34, 19–32.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Lade, P. V., Liggio, C. D., Jr., and Yamamuro, J. A. 1998. Effects of Non-Plastic Fines on
Minimum and Maximum Void Ratios of Sand. Geotechnical Testing Journal 21, 336–347.
Masin, D. 2019. Modelling of Soil Behaviour with Hypoplasticity: Another Approach to Soil
Constitutive Modelling, Springer Series in Geomechanics and Geoengineering. Springer
International Publishing, Cham, Switzerland.
Massarsch, K. R., and Fellenius, B. H. 2014. Ground vibrations from pile and sheet pile driving.
Part 1 Building Damage, in: Proceedings of the DFI-EFFC International Conference on
Piling and Deep Foundations. Stockholm, pp. 131–138.
Niemunis, A., and Herle, I. 1997. Hypoplastic Model for Cohesionless Soils with Elastic Strain
Range. Mechanics of Cohesive‐frictional Materials 2, 279–299.
Orozco-Herrera, J. E. 2021. Ground Movements and Vibrations Caused by Impact Pile Driving
of Prestressed Concrete Piles in Central Florida. University of Central Florida, Orlando,
Florida.
Orozco-Herrera, J. E., Turkel, B., Arboleda-Monsalve, L. G., Nam, B. H., and Jones, L. 2022.
Continuous Impact Pile Driving Modeling to Elucidate Settlement-PPV-Soil Density-Input
Energy Relationships, in: Geo-Congress 2022: Deep Foundations, Earth Retention, and
Underground Construction. Presented at the Geo-Congress 2022, American Society of Civil
Engineers, Charlotte, North Carolina, pp. 113–122.
Seed, H. B., and Idriss, I. M. 1970. Soil Moduli and Damping Factors for Dynamic Response
Analyses (No. EERC 70-10). EERC.
Turkel, B., Orozco-Herrera, J. E., Arboleda-Monsalve, L. G., Nam, B. H., and Jones, L. 2021.
Comparative Analysis of Pile Driving Numerical Modeling Approaches, in: International
Foundations Congress and Equipment Expo 2021. Presented at the International Foundations
Congress and Equipment Expo 2021, American Society of Civil Engineers, Dallas, TX, pp.
484–495.
von Wolffersdorff, P. A. 1996. A hypoplastic relation for granular materials with a predefined
limit state surface. Mechanics of Cohesive-Frictional Materials 1, 251–271.
Zapata-Medina, D. G., Vergara, C. Y., Vega-Posada, C. A., and Arboleda-Monsalve, L. G. 2019.
On the use of Fredlund gas–fluid compressibility relationship to model medium-dense gassy
sand behavior. Can. Geotech. J. 56, 1070–1079.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 455

Frost Susceptibility Evaluation of Clay and Sandy Soils

Mohammad Wasif Naqvi1; Md. Fyaz Sadiq2; Bora Cetin3; Micheal Uduebor4; and John Daniels5
1
Graduate Research Assistant, Dept. of Civil, Construction, and Environmental Engineering,
Michigan State Univ., East Lansing, MI. Email: naqvimo1@msu.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Graduate Research Assistant, Dept. of Civil, Construction, and Environmental Engineering,
Michigan State Univ., East Lansing, MI. Email: sadiqmd@msu.edu
3
Associate Professor, Dept. of Civil and Environmental Engineering, Michigan State Univ., East
Lansing, MI (corresponding author). Email: cetinbor@msu.edu
4
Graduate Research Assistant, Dept. of Civil and Environmental Engineering, UNC Charlotte,
Charlotte, NC. Email: muduebor@uncc.edu
5
Professor, Dept. of Civil and Environmental Engineering, UNC Charlotte, Charlotte, NC.
Email: jodaniel@uncc.edu

ABSTRACT

Frost action in soils causes a significant effect on the performance of roadways. This effect is
more pronounced in the regions that are experiencing seasonal subfreezing temperatures as the
soil undergoes multiple freeze-thaw cycles. Apart from the subfreezing temperature, the frost
action is also affected by the soil type as the void ratio and hydraulic conductivity of soils control
the presence and movement of water for the growth of ice lenses. Frost heave is mainly
attributed to silty soils, but significant frost heave can also occur in clay and sandy soils under
favorable environmental conditions. For the present study, frost heave and thaw settlement of
clayey and sandy soils, subjected to a one-dimensional freeze-thaw cycle, is investigated to
determine how the frost action varies with soil types. Soil specimens were subjected to ten
freeze-thaw cycles. Total heaving, heave rate, and water intake were measured as a function of
time during testing. The moisture content of the soils after ten freeze-thaw cycles was also
measured. The amount of pore water and external water supply affects the total heave during
freeze-thaw cycles. Therefore, the effect of moisture availability during the freeze-thaw cycles
was also investigated by comparing the results of specimens with or without an external water
supply. Results of the study suggested that significant frost heave occurred in both clay and
sandy soils. In addition, the application of ten freeze-thaw cycles provided a better estimation of
the total heave than that observed with two freeze-thaw cycles (typical/standard numbers of
freeze-thaw cycles). The maximum heave (40.9 mm) and heave rate (5.01 mm/day) were found
to be higher in clay soil. The presence of an external water supply contributed to the frost action,
and total heave was seven times higher in soils with an external water source. Soil with a free
water supply showed 1.1–1.7 times higher moisture content after ten cycles compared to the soils
with no external water supply. These results were used in estimating the frost heave potential of
soils in different environmental conditions.

Keywords: Frost susceptibility, frost heave, heave rate, heave ratio, water presence

INTRODUCTION

The frost action in soils affects the design, construction, and maintenance of geostructural
systems such as pavement challenging. The formation of ice lenses in soils can cause significant

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 456

heaving followed by a substantial strength reduction during thawing which may lead to a
systematic failure of the infrastructure. Current strategies for mitigating damages due to freeze-
thaw (F-T) is estimated yearly cost of over $2 billion (FHWA, 1999). For the frost action in
soils, three requirements are (1) frost susceptible soils, (2) freezing temperature, and (3) free
access to water (Chamberlain, 1981; Penner, 1959). The extent of damage caused by the F-T
depends on the previous three factors.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

It is accepted that some soils are more prone to frost heave action than others and are known
as frost susceptible soils, however, any type of soil can show frost heave if favorable conditions
are met (Bai et al., 2018; Naqvi et al., 2022; Sheng, 2021). The frost action in the soil is
dependent upon the suction and permeability of the soil (Carter & Bentley, 2016). Clay soils
have suction but very low permeability while sandy soils have high permeability but negligible
suction and water retention capacity. Therefore, intermediate particle size soils such as silts are
prone to be most frost susceptible. Soils in regions with seasonal freezing temperatures undergo
multiple F-T cycles causing failures in the infrastructure (Cetin et al., 2019). Studies have shown
that geotechnical properties of soils such as void ratio, porosity, permeability, plastic limit,
consolidation, resilient modulus, and shear strength are significantly affected by the number of
F-T cycles (Cui et al., 2014; Kumar & Soni, 2018; Qi et al., 2006; Rosa et al., 2016; Swan et al.,
2013). The ice formation in soils during freezing is due to the conversion of in situ pore water
and migrated water from an external source into ice. Studies have shown that in situ pore water
can only cause limited frost heaving due to limited water availability but when the water is freely
available frost heave is significantly high (Hermansson & Spencer Guthrie, 2005). The migration
of moisture toward the ice lens during freezing action is essential for continuous ice growth
(Dagli et al., 2018; Taber, 1930). In situ pore water contributes to heaving at the beginning of
freezing while the external migrated water is the main contributor once the freezing front is
ready (M. Zhang et al., 2017). The moisture content of the soil increases from the initial moisture
content after F-T cycles (Y. Zhang et al., 2016).
The present study aims to study the three governing factors (soil type, F-T cycles, water) for
frost actions. Two different soil types, clay, and sandy are subjected to a one-dimensional F-T
test. To understand the extent of F-T action, the soils are exposed to multiple F-T cycles. The
effect of moisture on the frost action is studied by testing soils with (open system) and without
water supply (closed system). Frost heave was measured to determine the heave rate and
maximum heave in soils subjected to multiple F-T cycles. In addition, the open system's water
intake was monitored. It was also assessed how much the moisture content within the specimen’s
top to bottom had changed during testing.

MATERIALS AND METHODS

The selected two soils are a low plasticity clay collected from Pottawattamie County, Iowa
(IA-PC) and silty sand collected from Boone County, North Carolina (NC-BO). These soils exist
in different climatic regions in the USA and are exposed to freeze-thaw conditions under
different environmental conditions. In the field, IA-PC experiences a wet, hard-freeze, spring
thaw while the NC-BO experience a wet, F-T cycling condition. Table 1 summarizes the index
properties of both soils. The IA-PC and NC-BO are classified as CL and SM/SC. The grain size
distribution curves of both soils are shown in Figure 1. Both of the soils contain a high
percentage of silt content and are therefore expected to exhibit adequate frost susceptibility
however the soils are primarily classified as clayey and sandy soils based on their index
properties. Figure 1 shows that the fines contents of IA-PC soil is higher than that of the NC-BO

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 457

soil. Both soils exist in F3 groups based on the frost susceptibility classification by the U.S.
Army Corps of Engineers (1965) (Table 1) and this indicates that both soils are frost susceptible
(ranging from low to very high).

Table 1. Summary of soil index properties and classifications of IA-PC and NC-BO soils
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Soil Properties IA-PC NC-BO

Specific Gravity, Gs (ASTM D854-14) 2.80 2.67

Liquid Limit, LL (%) (ASTM D4318-17) 37 38

Plasticity Index, PI (%) (ASTM D4318-17) 13 NP

Silt content (%) (75 μm–2 μm) 86.2 34.0

Clay content (%) (< 2 μm) 12.0 4.6

Optimum Moisture Content (%) (ASTM D698-12) 17.3 15.6

Max. Dry Unit Weight (kN/m3) (ASTM D698-12) 16.7 16.9

Saturated Hydraulic Conductivity (cm/s) 5.02e-08 6.34e-06

USCS Classification CL SM/SC

AASHTO Classification A-6 A-4

Frost Susceptibility Group+ F3 F3


*NP- Non-Plastic
+
Frost susceptibility classification by U.S. Army Corps of Engineers (1965) based on grain
size distribution.
F1-Low frost susceptibility; F4-Very high frost susceptibility

The frost heave testing of soil specimens was conducted using freezing-thawing test
equipment as shown in Figure 2 (per ASTM D5918). Two specimens of each soil type were
prepared. One specimen was subjected to a free external water supply (open system) while the
other sample was not subjected to any external water supply (closed system). To prepare the
specimens, the oven-dried soils were first crushed and passed through a 4 mm sieve. The
cylindrical samples were compacted with 33 blows in 6 layers using a standard proctor hammer.
The specimens were compacted at their optimum moisture content (OMC) (Table 1) and
maximum dry density. The specimen sizes were 14.6 cm in diameter and 15.2 cm in height. The
specimens were compacted within a latex membrane (to avoid leaks after inserting
thermocouples during testing) and 6 acrylic rings of 2.54 cm in height to restrain the sample
laterally. After compaction, the specimens were placed over a base plate that was connected to a
Marriot cylinder to act as a water source for the open system. For specimens with a closed
system, the base plate was not connected to the water supply. All the specimens were saturated
for 24 hours.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 458
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Grain size distribution of soils (ASTM D6913)

Figure 2. Frost heave testing assembly (Adopted from Mahedi, et al. 2020)

After saturation, the specimens were assembled for frost heave testing inside the chest freezer
which was programmed to maintain a temperature of 4º C. The sample along with the base plate
was placed on a bottom heat exchanger plate (also called a warm plate). Similarly, another heat
exchanger plate (also called a cold plate) was placed at top of the specimen to simulate the top-
down freezing conditions. These warm and cold plates were connected to a circulating bath for
controlling the temperature at the bottom and the top of the specimen. The fully programmable
circulating baths have a temperature range of -30º C to 200º C and can quickly change the
temperature of the circulating fluid. The temperature program of the two circulating baths is
provided in Figure 3. Since the specimens were subjected to freezing temperatures, the
circulating fluid was prepared by mixing water and ethylene glycol at a ratio of 1:1. A surcharge
weight of 2.25 kg was placed on each specimen. The specimens subjected to an open system
were connected to a Marriot bottle through a base plate and a constant head of 2.5 cm was
applied using a capillary tube during the experiment.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 459

The heaving of the specimen during the experiment was measured using lasers which had a
measuring range of 5 cm. A pressure transducer at the base of the Marriot bottle was installed to
measure the amount of water going inside the specimens during frost heave testing. The diameter
of the Marriot bottle was 7.62 cm. Thus, every unit cm water intake means 7.07 cm3 in volume.
The temperature of the specimen during experiments was measured with 8 T-type thermocouples
placed vertically at every 2.5 cm. The freezer was filled with packing peanuts and all the hoses
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

were thermally insulated to minimize any heat loss and avoid any external effects. The
specimens were subjected to 24-hour conditioning followed by 10 F-T cycles leading to a total
test duration of 21 days. The temperature program adopted for testing was similar to the ASTM
method as shown in Figure 3 except the F-T cycles were extended from two to ten. After the
experiment, the specimens were disassembled and sliced into 6 pieces to determine the
gravimetric moisture content of the specimens from top to bottom.

Figure 3. Temperature program in circulating bath for top and bottom heat exchange
plates

RESULTS AND DISCUSSION

Total Heaving, Heave Rate, Heave Ratio

Figure 4 presents the frost heave and thaw settlements under a number of F-T cycles. The
solid line represents the specimen with an open system while the dashed line shows the heave of
the specimens tested with a closed system. No heaving was observed during the first 24 hours
(conditioning period) as the temperature was above freezing temperature. All soils showed
heaving during freezing cycles and settlement during thawing cycles as expected. For an open
system, it is clearly evident that the specimens heave higher with an increase in the number of F-
T cycles. On the other hand, for a closed system, the magnitude of heave did not change with the
number of F-T cycles. This occurs due to the water available in the soil for the closed system
being limited thus its conversion to ice lenses during freezing temperatures is at minimal levels.
In addition, the absorbed and adsorbed water doesn’t convert into ice as per frozen fringe theory.
Therefore, the heave in soil with a closed system is expected to be lower and the heave stabilizes
from the first cycle for both soils in a closed system. The maximum heave of the closed system
specimens was significantly lower (5.73 mm for IA-PC) compared to those observed for the
open system (40.91 mm for IA-PC).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 460
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Heave and thaw of different soils during frost heave testing

For soils with the open system, the heave was considerably higher compared to the closed
system. This is largely in part due to the water migration induced by the temperature gradient
(Xu et al. 1999). A temperature gradient in the freezing soil means the development of water flux
in the direction of decreasing temperature which depends on the permeability of the frozen soil
and the suction force of the force fringe (Perfect & Williams, 1980) Since the water was freely
available, the specimens took external water that yields to large ice formations leading to a
greater magnitude of heaving (40.91 for IA-PC). This is also evident from the amount of water
intake during the experiment as discussed in the later section. The maximum heave was observed
in the first F-T cycle for NC-BO (open) which then reduced and finally stabilized at the 6th F-T
cycle. A similar trend was observed for IA-PC (open) where the heaving stabilized at the 6th
cycle. However, for IA-PC (open), the maximum heave was observed in the second F-T cycle.
This is caused by the lower initial hydraulic conductivity of IA-PC (Table 1) as a result of high
clay contents in IA-PC. After the first F-T cycle, the formation of large voids as well micro
fissuring may have lead to an increase in the hydraulic conductivity of IA-PC soil causing a
significant amount of heaving during the second F-T cycle. Then, similar to NC-BO, the
magnitude of heave decreased and stabilized at the 6th F-T cycle. The total heave of IA-PC in the
open (40.91 mm) was higher than that of NC-BO (31.54 mm). IA-PC soil had higher silt content
compared to NC-BO soil. As silt contributes to the heaving process the most, the total heave in
IA-PC soil was determined to be higher than that of NC-BO soil.
The total heaving during the experiment after two and ten F-T cycles is shown in Figure 5.
Current testing standards suggest testing two F-T cycles. However, soils in general especially in
harsher climatic regions are subjected to higher F-T cycles per year which may have a high
impact on the frost-heave behavior of soils. Figure 5 clearly shows this effect where the soils
subjected to 10 F-T cycles experienced up to 1.7 times higher frost heave amount than that of
those subjected to 2 F-T cycles. The heaving of both soils stabilizes after the 6th F-T cycle. These

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 461

results show that two F-T cycles may not be enough to gauge the frost heave potential of soil and
more F-T cycles may provide better insight into soil behavior during F-T conditions.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5. Maximum Heave after two and ten F-T cycles

The Frost heave ratio (𝜉) is defined as the proportion of the frost heave increment (𝛥ℎ) up to
frost depth (𝐻𝑓 ) over a certain period of time (equation 1). The frost depth in the soil is the depth
at which the freezing temperature is present and ice formation can occur. For the current
experiment, the temperature data from the thermocouple at the bottom of the specimen shows
that the freezing temperature was present at the bottom of the specimen. Therefore, frost depth is
considered as the height of the specimen in the present case and total heave after the 10th F-T
cycle is considered to calculate the frost heave increment. The frost heave ratio of the specimens
is shown in Figure 6a. IA-PC in an open system showed the maximum frost heave ratio (23.8 %)
followed by NC-BO (14.6) in an open system. On the other hand, the frost heave ratio for the
soils with the closed system was 1.9% and 2.2% for IA-PC and NC-BO, respectively.
𝛥ℎ
𝜉= ∗ 100% (1)
𝐻𝑓

(a) Frost Heave ratio (b) Maximum Heave rate

Figure 6. Frost Heave parameters

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 462

ASTM D5918 suggests the calculation of the heave rate from the first 8 hours of the freezing
period of the first and the second freezing cycle. For the present study, the larger magnitude from
the two cycles was considered as the heave rates of the soils. Figure 6b shows the heave rates of
both soils. Soils with open system IA-PC (5.01 mm/day) and NC-BO (4.11 mm/day) had higher
heave rates compared to closed system conditions for IA-PC (2.4 mm/day) and NC-BO (3.15
mm/day). Based on the heave rates of the tested soils, they can be classified as medium to highly
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

frost-susceptible soils.

Water Intake

Figure 7 shows the water intake by soils during the frost heave experiments along with the
corresponding heave data at the open system. The results corroborated with the frost heave data.
The general trend shows that more water enters the soils during freezing relative to thawing
because of the water flux toward specimens due to temperature gradient. This suggests that the
major contribution to heaving is due to external moisture sources. Excess water in fully saturated
soil retracts back into the Marriot bottle during thawing periods at a few thawing cycles. The
amount of water intake by sandy soil (NC-BO) was higher in the initial F-T cycles than that of
the clayey soil (IA-PC). Similar to the heave data, the water intake also stabilized after the 6th F-
T cycle and only limited water entered the specimens during the remaining F-T cycles. The total
water taken by the soils after 10 F-T cycles were similar.

Figure 7. Water intake by soils during testing

The moisture content of the specimens with depth after the F-T test is shown in Figure 8. The
moisture content of soils increased considerably from the initial optimum moisture content,
especially for the soils with open systems. The moisture contents of soils with an open system
were significantly higher than the ones with a closed system. It is worth noting that the moisture
content of the closed system specimen is higher than its OMC because of the saturation that is
applied before the frost heave test and the specimen primarily undergoes moisture redistribution

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 463

during the test. Since the freezing starts from the top, the moisture content increases towards the
top of the specimens due to the migration of the water toward the frozen fringe. Therefore, the
differences in the moisture contents between the bottom and top of IA-PC soil were 3.1% and
13.8% for a closed and open system, respectively. Similarly, this difference for NC-BO soil was
6.8% and 10.7%, respectively.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 8. Gravimetric moisture contents of soils after frost heave testing

CONCLUSIONS

Frost heave testing on sandy and clay soils was conducted. The long-term effects of F-T
cycles were investigated by applying 10 F-T cycles on the soils with (open system) and without
water supply (closed system). The heave was consistent after the first F-T cycle for both soils
when no water supply was provided. For soils with water supply (like shallow ground water
table), that maximum heave was observed in the first F-T cycle for sandy soil due to high
permeability while the maximum heave for clayey soil was observed during the second freezing
cycle due to the development of large voids and micro fissuring after the first F-T cycle causing
a significant increase in the permeability of IA-PC soil. The heave of both specimens
continuously increased up to the 6th F-T cycle and then stabilized. The total heave of IA-PC in
the open system was 7 times higher than that of the closed system. Higher heave ratios of
23.75% and 14.61% were found in the open system for IA-PC and NC-BO, respectively. The
differences in total heave after the second and the tenth cycle were significant for soil tested with
an external water supply. Therefore, multiple F-T cycles should be applied to better estimates the
heave potential of the soils. The water intake data during F-T cycles followed the heaving trends.
The moisture contents of the soils after ten F-T cycles differed noticeably between the open and
closed systems, and the open system soils showed considerable increases in moisture content
above their original optimal moisture contents. The movement of water towards the frozen fringe
caused the specimens' moisture levels to rise from the bottom to the top. This study will
contribute to the database of frost heave behavior of various soil types and assist in assessing the
extent of frost damage to pavements in various soil types subjected to various climatic conditions
and moisture availability.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 464

ACKNOWLEDGMENT

This research was sponsored by the National Science Foundation (Award #1928813) with
counterpart funding from the Iowa Highway Research Board.

REFERENCES
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Bai, R., Lai, Y., Zhang, M., and Gao, J. (2018). Water-vapor-heat behavior in a freezing
unsaturated coarse-grained soil with a closed top. Cold Regions Science and Technology,
155, 120–126.
Carter, M., and Bentley, S. P. (2016). Soil Properties and their Correlations, Second Edition.
Cetin, B., Satvati, S., Ashlock, J., and Jahren, C. (2019). Performance-Based Evaluation of Cost-
Effective Aggregate Options for Granular Roadways.
https://lib.dr.iastate.edu/intrans_techtransfer/123/.
Chamberlain, E. J. (1981). Frost susceptibility of soil Review of index tests.
Cui, Z. D., He, P. P., and Yang, W. H. (2014). Mechanical properties of a silty clay subjected to
freezing-thawing. Cold Regions Science and Technology, 98, 26–34.
Dagli, D., Zeinali, A., Gren, P., and Laue, J. (2018). Image analyses of frost heave mechanisms
based on freezing tests with free access to water. Cold Regions Science and Technology, 146,
187–198.
FHWA (Behavior. Federal Highway Administration). (1999). A Quarter Century of
Geotechnical Research, Chapter 4: Soil and Rock Behavior. Federal Highway Administration
(FHWA), Report Number: FHWA-RD-98-139.
Hermansson, Å., and Spencer Guthrie, W. (2005). Frost heave and water uptake rates in silty soil
subject to variable water table height during freezing. Cold Regions Science and Technology,
43(3), 128–139.
Kumar, A., and Soni, D. K. (2018). A Review on Freeze and Thaw Effects on Geotechnical
Parameters. Lecture Notes in Civil Engineering, 21 LNCE, 148–159.
Naqvi, M. W., Sadiq, M. F., Cetin, B., Uduebor, M., and Daniels, J. (2022). Investigating the
Frost Action in Soils. 257–267.
Penner, E. (1959). The mechanism of frost heaving in soils. Highway Research Board Bulletin,
225.
Perfect, E., and Williams, P. J. (1980). Thermally induced water migration in frozen soils. Cold
Regions Science and Technology, 3(2–3), 101–109.
Qi, J., Vermeer, P. A., and Cheng, G. (2006). A review of the influence of freeze-thaw cycles on
soil geotechnical properties. Permafrost and Periglacial Processes, 17(3), 245–252.
Rosa, M. G., Cetin, B., Edil, T. B., and Benson, C. H. (2016). Development of a Test Procedure
for Freeze-Thaw Durability of Geomaterials Stabilized With Fly Ash. Undefined, 39(6),
938–953.
Sheng, D. (2021). Frost susceptibility of soils-A confusing concept that can misguide
geotechnical design in cold regions. Sciences in Cold and Arid Regions, 13(2), 87–94.
Swan, C. W., Grant, A., and Kody, A. (2013). Characteristics of Chicago Blue Clay subjected to
a freeze-thaw cycle. ASTM Special Technical Publication, 1568 STP, 22–32.
Taber, S. (1930). The Mechanics of Frost Heaving. The Journal of Geology, 38(4), 303–317.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 465

Zhang, M., Zhang, X., Xu, X., Lu, J., Pei, W., and Xiao, Z. (2017). Water–heat migration and
frost-heave behavior of a saturated silty clay with a water supply. Experimental Heat
Transfer, 30(6), 517–529.
Zhang, Y., Johnson, A. E., and White, D. J. (2016). Laboratory freeze-thaw assessment of
cement, fly ash, and fiber stabilized pavement foundation materials. Cold Regions Science
and Technology, 122, 50–57.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 466

Measurement of Volumetric Deformation, Strain Localization, and Shear Band


Characterization during Triaxial Testing Using a Photogrammetry-Based Method

Sara Fayek, S.M.ASCE1; Xiong Zhang, Ph.D., P.E.2; Xiaolong Xia3; Qingqing Fu4;
and Jeffrey Cawlfield, Ph.D., P.E.5
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Ph.D. Candidate, Dept. of Civil, Architectural, and Environmental Engineering, Missouri Univ.
of Science and Technology, Rolla, MO. Email: fayeks@umsystem.edu
2
Professor, Dept. of Civil, Architectural, and Environmental Engineering, Missouri Univ. of
Science and Technology, Rolla, MO. Email: zhangxi@umsystem.edu
3
Graduate Student, Dept. of Civil, Architectural, and Environmental Engineering, Missouri Univ.
of Science and Technology, Rolla, MO. Email: xxrkq@umsystem.edu
4
Graduate Student, Dept. of Civil, Architectural, and Environmental Engineering, Missouri Univ.
of Science and Technology, Rolla, MO. Email: qfg2f@umsystem.edu
5
Professor, Dept. of Geosciences, Geological, and Petroleum Engineering, Missouri Univ. of
Science and Technology, Rolla, MO. Email: jdc@umsystem.edu

ABSTRACT

Triaxial testing has been routinely used as a standard laboratory test that allows correct
determination of soil characteristics. Previously the volumetric strain of the triaxial specimen
was considered to be uniformly distributed along with the specimen during the isotropic and
deviatoric loading. Although this assumption might hold true under isotropic loading, the effects
of restrained ends and disturbance during the procedures of specimen installation and testing can
cause nonuniform strains throughout the whole specimen. This paper investigates the effects of
specimen preparation and misalignment on the strain uniformity along with the soil specimen
during triaxial testing. A series of consolidated drained tests at several stress paths were
conducted on sand specimens. A photogrammetry-based method was applied at different stages
of specimen preparation and testing to provide a three-dimensional full-field deformation
measurement of the surface of the triaxial soil specimen. One commercial camera was used to
capture images for the triaxial specimen, and a developed application for data processing and
post-processing was utilized to ensure automatic and fast processing of the developed
photogrammetric-based method. The local displacement data provided by the photogrammetry-
based method enabled the evaluation of the strain localization and the volumetric strain
nonuniformity analysis at different heights along with the specimen. The triaxial test results
demonstrated that the soil specimen during triaxial testing has deformed nonuniformly in the
axial, radial, and circumferential directions. The plots of the strain localization precisely
presented the variation of local strains and the magnitude of deformation after the saturation
stage. These results prove the soil specimen volume is not constant during saturation, and
unavoidable disturbance had occurred during the specimen preparation steps and saturation. The
results proved that the specimen misalignment during triaxial testing leads to scattering in the
triaxial test results. Further discussion was presented about the shear band characterization
including shear band thickness, formation, and propagation.

Keywords: Triaxial Test, Photogrammetry, 3D full-field displacement, Strain Localization,


Shear Band.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 467

INTRODUCTION

Triaxial tests have been intensively used to characterize the behavior of both saturated and
unsaturated soils owing to their advantages of providing controlled well-defined boundaries,
uniform stresses within the soil, and drainage conditions (Fayek et al. 2022). During triaxial
testing, the soil volume and deformation are required to determine the stress-strain relationship
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

of soils. In addition, the accurate measurements of local and total deformation of soil specimens
contribute greatly to understanding the characteristics of nonlinearity and anisotropy behavior as
well as in the modeling and numerical simulation of soil mechanics (Alshibli et al. 2002; Desrues
and Viggiani 2004; Rechenmacher 2006). Quantification of shear band formation, growth, and
evolution is a crucial factor in the development of constitutive models of soils (Sachan and
Penumadu 2007). However, the triaxial soil specimen is subjected to many sources of
disturbance during preparation, installation, and saturation procedures besides the effect of
measurement techniques and equipment (Mulilis et al. 1977; Baldi et al. 1988; Santagata 1994;
and Scholey et al. 1995). Although previous research provided a useful guideline to understand
the effect of disturbance and error measurement on triaxial results, there is not yet a practical and
accurate methodology to quantify the effect of sample disturbance during triaxial testing.
In recent years with the improved development of image-based methods in triaxial testing,
the measurement of deformation and strain localization has gained increasing focus (Alshibli and
Sture 2000). The image-based methods can be classified in terms of their system designs and
principles into various categories including Digital image analysis (DIA), digital image
correlation (DIC), X-ray computed tomography (CT), and photogrammetric methods (Macari et
al. 1997; Gachet et al. 2007; Sachan and Penumadu 2007; Rechenmacher and Medina-Cetina
2007; Bhandari et al. 2012; Zhang et al. 2015). However, many of the previous methods only
considered part of the soil specimen and made assumptions about the soil specimen's shape and
deformation (Desrues and Viggiani 2004; Lin and Penumadu 2006; Rechenmacher 2006;
Bhandari et al. 2012). X-ray CT was able to detect the internal structure of the soil specimen by
interpreting the intensity data of the X-ray beam passes through the soil specimen. However, X-
ray CT is expensive to be implemented. For the first time, the photogrammetric method was able
to determine with cost-effectiveness the full-field deformation of soils during triaxial testing.
Photogrammetry is the science of reconstruction of 3D models from 2D images. Zhang et al.
(2015) developed a photogrammetry-based method to determine the specimen shape and
deformation at any stage of triaxial testing by using only one commercially available camera.
Multiple optical ray tracings and a least-square optimization technique were employed for
refraction correction at the air-acrylic cell and acrylic cell–water interfaces. Global and local
deformation measurements were performed for both saturated and unsaturated soils. However,
this method was not able to locate the top and bottom boundaries between the soil sample, the
top cap, and the bottom pedestal respectively since the top and bottom boundaries were covered
by a membrane during testing. As assumptions were made to determine the soil sample’s
boundaries, the accuracy of the soil volume measurements is questionable. To overcome this
limitation, Fayek et al. (2020) proposed a simple and rigorous photogrammetry-based technique
to determine the top and bottom boundaries of the soil specimens mathematically. The principle
is based on the fact that the distances from the peripheral coded targets to the surface plane of
each pedestal will remain unchanged, regardless of their locations and orientations during the
triaxial tests. So, by applying the method proposed by Fayek et al. (2020), the specimen ends and
the absolute soil volume were determined with high accuracy without making any assumptions

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 468

about the specimen’s shape or boundary. More improvement to the method was proposed by Xia
et al. (2021) by proposing a table method to make up missing points and allowed automatic and
fast processing of measurement targets on the soil specimen’s surface. With the implementation
of the photogrammetric method of Zhang et al. (2015), the absolute soil volume methodology of
Fayek et al. (2020), and the automatic processing presented in Xia et al. (2021), this paper
present a new approach that allows an automatic, fast, and accurate method to measure the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

volumetric deformation, strain localization, and shear band characterization during triaxial
testing.
This research allows a full-field deformation of soils during various stages of testing. A
series of consolidated drained tests at several stress paths were conducted on sand specimens. A
photogrammetry-based method was applied at distinct stages of specimen preparation and testing
to provide a three-dimensional (3D) full-field deformation measurement of the surface of the
triaxial soil specimen. The photogrammetric method proposed by Zhang et al. (2015) and the
improvements made by Fayek et al. (2020) and Xia et al. (2021) were applied with minor
modifications. One available commercial camera was used to capture images for the triaxial
specimen, and a developed application for data processing and post-processing was utilized to
ensure automatic and fast processing of the developed photogrammetric-based method. The local
displacement data provided by the photogrammetry-based method enabled the evaluation of the
strain localization and the volumetric strain nonuniformity analysis at different heights along
with the specimen. The comparison of the strain localization at distinct stages of specimen
preparation allowed the clarification of the potential sources of specimen disturbance. The
triaxial test results demonstrated that the soil specimen during triaxial testing is subjected to an
inevitable disturbance that caused deviation from its initial measured dimensions. The results
further discuss the formation of the shear band characterization including shear band thickness,
formation, and propagation. In addition, a conclusion on the effects of misalignment and
specimen preparation on the strain uniformity along with the soil specimen during triaxial testing
is made.

EXPERIMENTAL PROGRAM

Specimen Preparation and Triaxial Test Program

The soil used in triaxial testing is clean Ottawa 20-30 sand that conforms to ASTM
designated C778. The triaxial soil specimens of 72 mm in diameter were prepared using the
moist tamping method. Initially, the sand is oven-dried and sieved through a 0.5 mm screen.
Then, the dried soils are thoroughly mixed with a water content of 2 % water, then compacted in
6 layers. The soil specimen can reach the desired density by adjusting the number of tamps per
layer, the force per tamp per layer, and the amount of water added during mixing. The under-
compaction method, which simulates a relatively homogeneous soil condition, was used to
prevent excessive densification of the lowest layers (Ladd 1978). Each subsequent layer received
less percentage of under-compaction, where the tamping energy is gradually augmented from the
bottom to the top layer so that uniform density can be obtained throughout the specimen layers.
In addition, the top of each compacted layer was scarified before the addition of material for the
next layer to ensure good contact with the subsequent soil layer. This process was repeated until
a full height of 144 mm was reached for the soil specimen of a relative density of 70%. Then the
mold was removed, and the membrane was sealed at the top cap and base pedestal with rubber
O-rings. After that, the triaxial chamber was assembled and filled with water.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 469

Initially, the soil specimen was percolated with carbon dioxide gas for 45 minutes to push the
air up through the top drainage line. Then, the soil sample was slowly flushed with de-aired
water to push most of the carbon dioxide out. Any carbon dioxide that remained in the specimen
can be dissolved in the intruding water, and subsequently, fill the voids in the soil specimen. To
ensure that the specimen is fully saturated, the back-pressure saturation procedure was applied
until the measurement of the B-value (B-pore pressure parameter) reached at least 0.98. Other
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

specimens were prepared with the same water content and similar preparation procedure but
tested at different stress path. For this paper, the results are presented for only one specimen. The
specimen was isotropically compressed to 200 kPa, then unloaded to 50 kPa, and later reloaded
to 300 kPa. The specimen was sheared by advancing the axial piston downward, with a strain
rate of 0.5% per hour, under a constant confining pressure of 300 kPa until failure. The volume
of water exchange from the sample, the axial load-deformation, and the volume change were
recorded.

Photogrammetry-Based Method: Implementation and System Setup

The photogrammetric method proposed by Zhang et al. (2015) besides the improvement
made by Fayek et al. (2020) and Xia et al. (2021) were applied in this research with slight
modification. The conventional triaxial test apparatus for saturated soils, pressure sensors, digital
controllers, and a close-range photogrammetry-based system were used to measure the
displacement of soil specimens after each preparatory, compression, and shearing stage (Figure
1). One minor modification was applied to the top cap and bottom pedestal by partially
engraving their surface to place porous stones. To prevent the block view caused by the drainage
lines in the triaxial chamber, the drainage lines connecting the top pedestals with the triaxial cell
base were relocated to be connected with the triaxial cell top.

ell pressure transducer

Measurement argets

oil specimen
ore pressure transducer

amera

Figure 1. Triaxial testing system presenting a deformed soil specimen during the shearing
phase.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 470

The close-range photogrammetry system used during triaxial testing consisted of the
following:
(1) A commercially available digital camera (Nikon D7000) with a 50-mm fixed focal length
lens (AF-S Nikkor 50mm f/1.4 G) to capture the images. The camera has an image sensor
with a resolution of 16.2 million pixels (4928 H: 3264 V). The camera was fully
calibrated before initiating the triaxial testing to acquire accurate and reliable 3D metric
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

information from the captured images.


(2) A photogrammetry commercial software for the camera calibration process.
(3) A developed MATLAB program for data post-processing. The top and bottom
boundaries between the soil specimen, and the top cap and bottom pedestal, were
determined using the developed program. Also, the specimen surface interpolation,
volume calculation, displacement, and strain were automatically calculated.
(4) A latex membrane with a total number of 816 measurement targets consisting of 204
black-ringed automatically detected (RAD) coded targets and 612 black dots printed on
its surface.
Additional RAD targets were posted on the outside surface of the acrylic chamber, including
two circles and three vertical stripes to determine the cell wall deflection and the cell shape and
location at different confining pressure during testing. Also, RAD targets were posted on the
triaxial cell tie rods and load frame to provide overlapping points between the captured images
and enable the establishment of a global coordinate system for the photogrammetry-based
analysis. In addition to that, measurement targets were placed on the top surface and periphery of
both the top cap and bottom pedestal. Using a digital caliper, several measurements between
RAD targets were acquired to define the scale as an input in the MATLAB app. The procedure
of the photogrammetry-based method for determining the deformation of the soil specimen
during triaxial testing can be summarized in the following steps:
• Camera calibration: To achieve a high level of measurement accuracy in the
photogrammetric analysis, the camera calibration process is required. Since commercially
available cameras often use multiple lenses, slight bending of the light rays either
outward or inward is caused. Thus, the camera should be calibrated to get precise and
reliable 3D reconstruction of the soil specimen. In this study, the calibration process was
performed using a single calibration sheet. The calibration is performed by capturing
using the camera around 12 images of the calibrated sheet. By analyzing these images,
the intrinsic and extrinsic parameters were calculated.
• Photographing: The calibrated camera was used to capture images around the triaxial
apparatus at a different stage of triaxial testing. First, images were captured for the top
cap and bottom pedestals independently which can be used to back-calculate the
specimen’s boundaries with the top cap and bottom pedestal. Additional pictures were
taken for the assembled model during triaxial testing. More details about the process of
back calculating the top and bottom boundaries of the soil specimen are presented in
Fayek et al. (2020).
• Photogrammetric analysis: The camera calibration parameters were imported in the
MATLAB app. Then the camera orientations, location, and the 3D coordinates of the
targets were determined by the photogrammetric analysis. During this process, the acrylic
cell shape and deformation were acquired for the ray-tracing technique and the least-
square optimization technique. More details about this procedure can be found in Zhang
et al. (2015).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 471

• Post-Processing: After acquiring the 3D coordinates of the targets on the surface of the
soil specimen, the top and bottom boundaries between the soil specimen, the top cap, and
the bottom pedestal were back-calculated (Fayek et al. 2022b). The displacement and
strain localization calculations were determined using the method described by Lin and
Penumadu (2006) as presented in the following section.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

STRAIN LOCALIZATION: INTERPOLATION AND CALCULATIONS

The photogrammetry-based method is used to evaluate the deformation of the soil specimen
during triaxial testing. The 3D coordinates of the 816 discrete measurement points on the soil
surface were determined at desired time intervals (Figure 2.a). Then, the 3D coordinates at two
different time intervals were compared, for example comparing the 3D coordinates after
saturation to those at the initial condition in the air. The displacement of each measurement
target during each time interval was calculated. The strain localization calculations were adopted
from Lin and Penumadu (2006). A small 3D block was separated from the specimen, as shown in
Figure 2.b, with four targets on the outside surface whose 3D coordinates are measured using the
photogrammetric method. The depth of the block is considered to be around 10 mm. This block
is further divided into five irregular tetrahedrons where each point of the tetrahedron has three
displacement components including vertical displacement, u1; radial displacement, u2; and
circumferential displacement, u3.

Figure 2. (a) Targets printed on the membrane surface; (b) The 3D cube and its divided
five four-surface elements (Modified from Lin and Penumadu (2006)).

The displacement of any point within the element was assumed to be a linear function of the
coordinates z, r, and θ as per Equation 1 (Lin and Penumadu 2006) as follows:

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 472

u m =a m0 + a m1z + a m2 r + a m3θ (1)

where amn (m=1 to 3, n=0 to 3) are constant to be determined.


Using Lin and Penumadu (2006) suggested procedure, the local strain components can be
solved using Equation (2). It is worth noting that Equation 2 was corrected using finite element
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

theory as follows:

 u1 
 z 
 
 u2 
 z   r 
   u2

 r   

      r 

 =
  1  u 
u3 (2)
 r   
u
3
+ − 3 
 rz   2  r r  r 
   1  u u2  
  z 
   
1
+  
 2  r z  
 1 u u3  
  

2
+ 
 2  r 
 z  

The displacement components of points 5, 6, 7, and 8 can be calculated considering a linear


relationship with points 1, 2, 3, and 4. The vertical, radial, and circumferential displacements of
points 5, 6, 7, and 8 are presented in Equations 3, 4, and 5 respectively as follows:

u15 = u11; u16 = u12 ; u17 = u13 ; u18 = u14 (3)

r5 r r r
u25 = u21  ; u26 = u22  6 ; u27 = u23  7 ; u28 = u24  8 (4)
r1 r2 r3 r4

r5 r r r
u35 = u31  ; u36 = u32  6 ; u37 = u33  7 ; u38 = u34  8 (5)
r1 r2 r3 r4

RESULTS

Using the 3D coordinates obtained from the photogrammetry-based method and the
formulation presented in the previous section, the displacement and strain components were
calculated. To facilitate the visualization of the uniformity of deformation and the strain
localization of the soil specimen, contour plots are used to illustrate the 3D surface displacement
and strain localization. The points that share the same value of displacement or strain are
connected by contours with different colors that represent the value corresponding to the
displacement or strain. To generate a full-field continuous deformation of the specimen, a
numerical interpolation technique was used between the discrete measurement points. Figure 3

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 473

shows the 3D full-field displacement and strain localization of the soil specimen after saturation.
The intensity of the color on the contour plots represents the magnitude of the corresponding
displacement or strain. Following the convention used in geotechnical engineering, a negative
sign for displacement and strain represents extension, and a positive sign represents compression.
It can be noticed that the soil specimen has deformed nonuniformly after saturation with a
maximum of 2 mm, 1 mm, and 0.04⁰ in the axial, radial, and circumferential directions,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

respectively. Strains values vary between 0 and 0.08, -0.04 and 0.04, and -0.04 and 0.04 in the
axial, radial, and circumferential directions, respectively. The specimen response is non-uniform
in the axial, radial, and circumferential strain, and they are positive in the left half of the
specimen. Negative displacement and strain were noticed on the right side of the soil specimen.
As the specimen was assumed to have negligible disturbance during the specimen preparation
and saturation stages, theoretically, no strain localization was expected. However, the plots of the
strain localization accurately present the variation of local strains and the magnitude of
deformation after the saturation stage. These results prove the soil specimen volume is not
constant during saturation and unavoidable disturbance had occurred during the specimen
preparation steps and saturation.

Figure 3. 3D full-field displacement and strain localization: (a) axial displacement, (b)
radial displacement, (c) circumferential displacement, (d) axial strain, (e) radial strain, (f)
circumferential strain, (g) ԐRZ (h) ԐRϴ (e) ԐϴZ.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 474

Similarly, the circumferential strain at the end of the shearing stage was calculated and
presented as shown in Figure 4. This result is the best presentation of the trend of development of
the shear band. The angle and thickness of the shear band were derived from the contours on the
soil specimen. The shear band with clear inclination starts developing at 3% global vertical strain
and becomes more significant with the increase of global vertical strain. The inclination and the
thickness of the shear band, denoted by θ and t respectively, were determined by the dense color
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

of the contour lines. In addition, it can be noticed from Figure 4 that the top edge of the soil
specimen is not horizontal at the end of the shearing stage. The tilting of the top cap or soil
specimen can be attributed to many factors as described by Fayek et al. (2022c). The excess
deformation or formation of the shear band can be correlated to the misalignment of the soil
specimen. This result is also related to the findings of Fayek et al. (2022c) where the misaligned
specimen demonstrated higher deformation compared to the idealized aligned specimen case.
This proves that the specimen misalignment during triaxial testing leads to scattering in the
triaxial test results. Specimen misalignment alters the strain distribution uniformity in the triaxial
specimen and resulted in higher strain values. This influence becomes larger and larger with
increasing strains and these nonuniformities can significantly influence the estimation of the soil
shear strength properties. Discussions regarding vertical displacement, tilting, and eccentricity
can be found in (Fayek et al. 2022a; Fayek et al. 2022c).

Figure 4. 3D full-field circumferential strain showing the shear band inclination and
propagation.

CONCLUSION

A noncontact photogrammetry-based method was used to reconstruct an accurate 3D full-


field model of the soil specimen at any stage of triaxial testing. Further analysis was performed
to determine the specimen’s ends and the absolute soil volume. In addition, the strain localization
was calculated using the finite element theory. Based on the displacement and strain localization
results, the soil was noticed to be deformed non-uniformly after the saturation process. This
confirms that the saturation process can disturb the soil specimen and that the volume of soil
during the preparatory and saturation stage is not constant. During deviatoric loading, the soil has
further deformed, and a shear band was formed. The calculation of the shear band inclination can
be easily derived from the color of contour lines. Based on the measurement results on the
triaxial soil specimen, it could be concluded that the photogrammetry-based method proposed is

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 475

a powerful tool for an in-depth understanding of soil behavior during triaxial testing. More
details on the comprehensive evaluation of important aspects of triaxial testing using the
photogrammetry-based method can be found in Fayek et al. (2023).

REFERENCES
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Alshibli, K., and Sture, S. (2000). “Shear band formation in plane strainexperiments of sands.” J.
Geotech. Geoenviron. Eng.,126(6), 495–503.
Alshibli, K. A., Sture, S., Costes, N. C., Frank, M. L., Lankton, M. R., Batiste, S. N., and
Swanson, R. A. (2000). “Assessment of localized deformations insand using x-ray computed
tomography.” Geotechn. Test. J., 23(3), 274–299.
Baldi, G., Hight, D. W., and Thomas, G. E. (1988). A reevaluation of conventional triaxial test
methods.ASTM STP977, ASTM, West Conshohoeken, Pa., 219–263.
Bhandari, A. R., Powrie, W., and Harkness, R. M. (2012). “A digital image-based deformation
measurement system for triaxial tests.” Geotechn. Test. J., 35(2), 209-226.
Desrues, J., and Viggiani, G. (2004). “Strain localization in sand: an overview of the
experimental results obtained in Grenoble using stereophotogrammetry.” Int. J. Numer. Anal.
Methods Geomech., 28(4), 279-321.
Fayek, S., Xia, X., Li, L., and Zhang, X. (2020). “Photogrammetry-based method to determine
the absolute volume of soil specimen during triaxial testing.” Transportation Research
Record, 2674(8), Transportation Research Board, Washington, D.C., 206-218.
Fayek, S., Xia, X., and Zhang, X. (2022a). “Consideration of One Camera Photogrammetry-
Based Method to Reevaluate Some Aspects of Conventional Triaxial Testing.” In Geo-
Congress 2022, 141-151.
Fayek, S., Zhang, X., Galinmoghadam, J., and Cawlfield, J. (2022b). “Point Density for Soil
Specimen Volume Measurements in Image-based Methods During Triaxial Testing.” Acta
Geotech. (Tentatively accepted).
Fayek, S., Zhang, X., Galinmoghadam, J., and Xia, X. (2022c). “Evaluating the Effects of
Specimen Misalignment during Triaxial Testing Using a Photogrammetry-Based Method,”
Geotechn. Test. J., (Under Review).
Fayek, S., Zhang, X., and Xia, P. (2023). “Comprehensive Evaluation of Important Aspects of
Triaxial Testing Using a Photogrammetry-based Method,” Transportation Research Record,
(Tentatively accepted).
Gachet, P., Geiser, F., Laloui, L., and Vulliet, L. (2007). “Automated digital image processing
for volume change measurement in triaxial cells.” Geotechn. Test. J., 30(2), 98-103.
Ladd, R. S. (1978). “Preparing test specimens using undercompaction.” Geotechn. Test. J., 1(1),
16-23.
Lin, H., and Penumadu, D. (2006). “Strain localization in combined axial-torsional testing on
kaolin clay.” J. Eng. Mech., 132(5), 555-564.
Macari, E. J., Parker, J. K., and Costes, N. C. (1997). “Measurement of volume changes in
triaxial tests using digital imaging techniques.” Geotechn. Test. J., 20(1), 103-109.
Mulilis, J. P., Seed, H. B., Chan, C. K., Mitchell, J. K., and Arulanandan, K. (1977). “Effects of
sample preparation on sand liquefaction.” J. Geotech. Eng. Division, 103(2), 91-108.
Rechenmacher, A. L. (2006). “Grain-scale processes governing shear band initiation and
evolution in sands.” J. Mech. Phys. Solids, 54(1), 22-45.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 476

Rechenmacher, A. L., and Medina-Cetina, Z. (2007). “Calibration of soil constitutive models


with spatially varying parameters.” J. Geotech. Geoenviron Eng., 133(12), 1567-1576.
Sachan, A., and Penumadu, D. (2007). “Strain localization in solid cylindrical clay specimens
using digital image analysis (DIA) technique.” Soils Found., 47(1), 67-78.
Santagata, M. C. (1994). Simulation of sampling disturbance in soft clays using triaxial element
tests (Doctoral dissertation, Massachusetts Institute of Technology).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Scholey, G. K., Frost, J. D., Presti, D. L., and Jamiolkowski, M. (1995). “A review of
instrumentation for measuring small strains during triaxial testing of soil specimens.”
Geotechn. Test. J., 18(2), 137-156.
Xia, X., Zhang, X., Fayek, S., and Yin, Z. (2021). “A table method for coded target decoding
with application to 3-D reconstruction of soil specimens during triaxial testing.” Acta
Geotech., 16(12), 3779-3791.
Zhang, X., Li, L., Chen, G., and Lytton, R. (2015). “A photogrammetry-based method to
measure total and local volume changes of unsaturated soils during triaxial testing.” Acta
Geotech., 10(1), 55-82.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 477

Small-Strain Behavior and Stress Path Rotation Angle Effects of Hawthorn Group Sands
in Central Florida

A. J. Aparicio-Ortube1; Luis G. Arboleda-Monsalve2; David G. Zapata-Medina3;


and Larry Jones4
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Doctoral Student, Departamento de Ingeniería Civil, Universidad Nacional de Colombia, Sede
Medellín, Medellín, ANT, Colombia. Email: aaparicioo@unal.edu.co
2
Assistant Professor, Dept. of Civil, Environmental, and Construction Engineering, Univ. of
Central Florida, Orlando, FL. Email: luis.arboleda@ucf.edu
3
Associate Professor, Departamento de Ingeniería Civil, Universidad Nacional de Colombia,
Sede Medellín, Medellín, ANT, Colombia. Email: dgzapata@unal.edu.co
4
State Geotechnical Engineer, Florida Dept. of Transportation, Tallahassee, FL.
Email: larry.jones@dot.state.fl.us

ABSTRACT

The small-strain behavior and effects of stress path rotation angle of Hawthorn Group (HG)
sands in Central Florida are examined in this paper. A series of directional-stress-probe tests
were conducted on high-quality HG sand samples extracted from the deep foundation test site
(DFTS) at the University of Central Florida (UCF). The small-strain response was studied under
undrained and drained triaxial conditions following compression and extension stress paths using
internal instrumentation, bender elements (BE), and Hall effect (HE) displacement transducers.
For the entire testing program, the stress history of the material was reproduced during the
reconsolidation stage in order to isolate the effects of stress path rotation angle. The discussion is
centered on the small-strain behavior with emphasis on shear, bulk, and cross-coupling moduli
degradation. The results showed that the small-strain stiffness behavior of this material is
affected by the stress path rotation angle.

INTRODUCTION

Investigation on the stiffness response of soils is required in many geotechnical engineering


applications to accurately predict ground deformations under expected working conditions,
especially when serviceability criteria are of interest rather than ultimate limit resistance
considerations. Since Burland (1989) reported that working strain levels in soils were less than
0.1% in well-designed geo-structures, a lot of effort has been devoted in studying the soil
response at small strains. In triaxial testing, the measurement of soil response in the range of very
small strains (<0.001%) and small strains (0.001% to 0.1%) requires special devices since the
standard external equipment coupled to the triaxial apparatus to measure soil specimen
deformations is affected by seating, alignment, bedding, and compliance errors (Scholey et al.
1995). Typically, small-size devices, such as linear variable differential transformers (LVDT)
and Hall Effect (HE) displacement transducers, mounted directly on the soil specimen within the
triaxial cell are employed to measure soil response in the range of small strains while Bender
Elements (BE) are used to measure stiffness response at very small strains via the interpretation
of wave propagation signals. On-specimen instrumentation has been employed in various
laboratory testing programs (e.g., Atkinson et al. 1990; Cho and Finno 2010; Finno and Cho
2011; Finno and Kim 2012; Zapata-Medina et al. 2014; Aparicio-Ortube et al. 2021), and the

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 478

results have shown that the stiffness soil response is highly non-linear, decays with strain level,
and is influenced by the stress path rotation angle, 𝜃, defined as the angle between the current
stress direction and its previous stress path. Stiffer soil responses were observed when the stress
path rotation angle approaches 180°, i.e., when the direction of the applied stress approaches the
complete reversal.
Research on the mechanical behavior of sands is mostly conducted on reconstituted
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

specimens in laboratory testing programs. A number of studies are found in the literature where
the effects of variables related to the physical nature and state of sands (e.g., grain-size
distribution, grain shape, mineral composition, fines content, stress state, fabric, and relative
density) on the stress-strain-strength response were evaluated. Despite the advantages of
employing reconstitution techniques in laboratory testing, comparative studies including high-
quality sand samples have shown that the reconstitution technique influences the stress-strain-
strength response and may not reflect the mechanical response of the soil mass at in situ
conditions (Vaid et al. 1999; Høeg et al. 2000; Yamamuro and Wood 2004; Huang and Huang
2007). Extraction of high-quality coarse-grained soil samples from the test site is a complex task,
time-consuming and expensive. However, if the sand material is below the water table, presents
clay minerals in its structure, and was subjected to an aging process, the interparticle bonding
developed during its geologic history and capillary forces allow the easy extraction of high-
quality sand samples and do not require the use of complex high-quality sampling methods, as in
the case of Hawthorn Group (HG) sands in Central Florida. The study conducted by Aparicio-
Ortube et al. (2022) on the compressibility characteristics of HG soils has shown that the
extraction of high-quality sand samples via thin-walled Shelby tubes is possible, and
representative soil mechanical parameters of the in situ conditions can be obtained when soil
specimens are tested under triaxial conditions.
The HG has been widely studied from a geological approach because of its economic and
hydrologic importance in the region. Research on its mechanical behavior is scarce in the
technical literature although the attention of researchers to this geologic unit has increased in the
last decades due to the severe damage to urban infrastructures by sinkhole activity (e.g., Shamet
et al. 2018; Kim et al. 2020; Nam and Shamet 2020; Soliman et al. 2020). Between 2006 and
2010, $1.4 billion was estimated based on 24,761 insurance claims for sinkhole damage in the
state of Florida (Weary 2015). The assessment of ground deformations in sinkhole-prone areas
requires reliable predictions, emphasizing the need to investigate the small-strain behavior of the
soils in this region. In this paper, the results and analyses of a series of directional-stress-probe
tests conducted to examine the small-strain behavior and effects of stress path rotation angle of
HG sands in Central Florida are presented. High-quality sand samples extracted from the Deep
Foundation Test Site (DFTS) at the University of Central Florida (UCF) were tested at in situ 𝐾0 -
conditions following compression and extension stress paths under undrained and drained triaxial
conditions. On-specimen instrumentation, BE and HE displacement transducers, was employed to
obtain the shear, bulk, and cross-coupling moduli response at different strain levels. The effects of
stress path rotation angle were considered when evaluating the small-strain stiffness responses.

EXPERIMENTAL PROGRAM

Soil Samples

Laboratory testing was conducted on HG sand samples obtained during soil exploration of
the DFTS at the UCF. The material was extracted from the northeast corner at elevations from

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 479

2.1 m to 4.4 m above mean sea level (MSL) employing six thin-walled Shelby tubes of 30 in
(760 mm) long with an external diameter of 3 in (76 mm). Extraction of the HG sand samples
was possible without employing complex high-quality sampling methods since the material was
below the water table, presented clay minerals in its structure, and was subjected to an aging
process. The sands of the HG are structured soils, and the water table was found at an elevation
of 16.1 m above MSL. In sandy soils where the interparticle bonding is absent or slightly
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

developed, high-quality samples can be obtained by freezing the groundwater in order to fix the
soil particles in the frozen ground (ground freezing method). The HG is a geologic unit formed
during the Miocene epoch and composed of clayey sands to silty clays and relatively pure clays,
silt-sized dolomite cemented sands, and variably quartz sandy dolostones (Scott et al. 1981; Scott
1988, 2001; Green et al. 2015; Upchurch et al. 2019). Post-depositional processes arising from
sea-level fluctuations, precipitation of cementing agents, and aging are the principal causes of the
small degree of overconsolidation (OCR≈2.5) of HG soils (Kenney 1964; Bjerrum 1967). Details
on the soil exploration program, stratigraphic sequence, geology, compressibility response, soil
structure, and field tests performed at the DFTS can be found in Aparicio-Ortube et al. (2022).
The extracted HG samples from the northeast corner of the DFTS were classified as silty
sands (SM) according to the unified soil classification system (USCS). However, gradation
results showed that the fine fraction is composed of silts and clays varying from 1.7% to 9.8%
and from 4.6% to 29.3%, respectively, indicating that this soil is a clayey sand material in
concordance with the geological description. The natural water content was approximately 30%
for the entire tests and the liquid and plastic limits varied from 0% to 22% and from 0% to 17%,
respectively. A series of oedometer tests conducted on this material and interpreted via the strain
energy method (Becker et al. 1987) indicate that the preconsolidation pressure, 𝜎′p , is
approximately 230 kPa (Aparicio-Ortube et al. 2022).

Laboratory Equipment, Specimen Preparation, and Testing procedures

The stress-probe tests were performed in an advance dynamic triaxial testing system
manufactured by GDS Instruments©. Figure 1(a) illustrates the triaxial apparatus with the
external and internal instrumentation used in this experimental program. The external
instrumentation consisting of two standard digital pressure/volume controllers and a pore
pressure transducer was employed to control pressures and volume changes, and measure pore
water pressures at the base of the specimens, respectively. The internal instrumentation
consisting of a submersible load cell, a set of HE displacement transducers, and a set of vertical
BE was used to measure axial loads, local displacements, and shear wave velocities, 𝑉s ,
respectively. The two axial HE displacement transducers are composed of a linear output HE
semiconductor housed in a metallic container and a spring-mounted pendulum with a magnet
assembly while the radial HE displacement transducer of a caliper with a magnet assembly and a
HE semiconductor placed at the opening of the caliper. Pads, pins, and adhesive were used to
mount the set of HE displacement transducers in the middle third of the specimens. These
transducers have a linear range of ±3.0 mm around the electrical zero and are capable of
resolving shear and volumetric strains as low as 0.02% and 0.04%, respectively, for the
specimens tested herein. The set of BE was installed in the top cap and base pedestal and aligned
such that vertically propagated waves were transmitted.
Soil samples were extracted from the thin-walled Shelby tubes, and specimens of 150 mm in
height and 70 mm in diameter, yielding a height-to-diameter ratio of 2.1, were hand-trimmed.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 480

Prior to saturation, the effective residual stress, 𝑝′r , was estimated via a pore water pressure
measurement technique suggested by Ladd and Lambe (1964). The specimens were back-
pressure saturated at 𝑝′r by simultaneously increasing the cell and back pressures to minimize
disturbance induced by swelling (Cho et al. 2007). Saturation was completed when the pore
pressure Parameter B was larger than 0.95. Figure 1(b) illustrates the stress paths followed
during reconsolidation and stress probes imposed at in situ stress state with dashed and solid
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

lines, respectively, and Table 1 describes the stress-probe tests performed in this experimental
program. The modified SHANSEP approach (Zapata-Medina et al. 2014) was employed to
reconsolidate the specimens by reproducing the stress history of the material. In this approach,
the soil specimen is loaded to 𝜎′p and unloaded to 𝜎′v0 under 𝐾0 -conditions (i.e., zero lateral
strain). For the entire testing program, the stress paths followed during reconsolidation were the
same and performed under stress-controlled conditions at a constant radial stress rate of 0.05
kPa/min. The axial stresses were continuously adjusted to minimize radial strains and assure 𝐾0 -
conditions. With this approach, the effects of stress path rotation angle on the subsequent
mechanical response were isolated. Once in situ stress state was reached, a drained creep period
under constant stress conditions was initiated to bring the specimens to rest prior to stress-probe
testing and lasted until the creep axial strain rate was less than 0.0020%/h to avoid effects of
continued deformations on the measurement of initial responses during the subsequent stage
(Santagata et al. 2005). For the specimens tested herein, the creep stage lasted approximately 2
hours. Throughout reconsolidation and creep, drainage at the base of specimens was prevented
such that excess pore water pressures could be measured. Also, BE tests were performed
inducing single-pulse sinusoidal input signals with an amplitude of 14 V and a frequency of 5
kHz. Electrical noise and near-field effects were minimized since ten BE signals were stacked
per each BE test and the ratio of wave travel distance to wavelength was larger than two (Kim et
al. 2015). A series of directional-stress-probe tests were conducted at in situ 𝐾0 -conditions under
undrained and drained triaxial conditions following compression and extension paths. The
undrained stress probes were conducted under strain-controlled conditions at a constant strain
rate of 0.2%/h until capturing the small-strain response, then a constant strain rate of 0.5%/h was
imposed, whereas the drained stress probes were performed under stress-controlled conditions
with a maximum stress rate of 3 kPa/h.

Figure 1. (a) Schematic diagram of the triaxial apparatus equipped with external and
internal instrumentation at the UCF geotechnical laboratory, and (b) stress paths followed
during 𝑲𝟎 -reconsolidation and stress probes on HG sands.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 481

Table 1. Stress-probe tests on HG sands.

Specimen Shelby Elevation 𝝈′𝐯𝟎 Stress


Description
ID tube [m MSL] [kPa] Probe
SP-1 B1-ST2 3.7 115 𝑈𝑇𝑋𝐶 Undrained triaxial compression
𝑈𝑅𝑇𝑋𝐸
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

SP-2 B1-ST2 3.7 115 Undrained reduced triaxial extension


Constant mean normal stress
SP-3 B1-ST3 2.9 120 𝐶𝑀𝑆
compression
SP-4 B1-ST3 2.9 120 𝐶𝑀𝑆𝐸 Constant mean normal stress extension
SP-5 B2-ST5 4.4 110 𝐶𝑄𝐿 Constant shear loading
SP-6 B2-ST6 3.7 115 𝐶𝑄𝑈 Constant shear unloading
Note: 𝜎′v0: in situ effective vertical stress.

Data Interpretation and Terminology

The following definitions are used for the effective mean normal stress, 𝑝′, deviator stress, 𝑞,
volumetric strain, 𝜀vol , and shear strain, 𝜀sh , under axisymmetric conditions:

𝑝′ = (𝜎 ′ a + 2𝜎 ′ r )/3; 𝑞 = 𝜎 ′ a − 𝜎 ′ r (1)

𝜀vol = 𝜀a + 2𝜀r ; 𝜀sh = 2(𝜀a − 𝜀r )/3 (2)

where 𝜎′a , 𝜎′r , 𝜀a , and 𝜀r are the axial and radial effective stresses and axial and radial strains,
respectively. The shear, 𝐺sec , bulk, 𝐾sec , and cross-coupling, 𝐽s sec and 𝐽v sec, secant moduli are
defined as:

𝐺sec = Δ𝑞/3Δ𝜀sh when Δ𝑝′ = 0 (𝐶𝑀𝑆 and 𝐶𝑀𝑆𝐸) (3)

𝐾sec = Δ𝑝′/Δ𝜀vol when Δ𝑞 = 0 (𝐶𝑄𝐿 and 𝐶𝑄𝑈) (4)

𝐽s sec = Δ𝑝′/Δ𝜀sh when Δ𝑞 = 0 (𝐶𝑄𝐿 and 𝐶𝑄𝑈) (5)

𝐽v sec = Δ𝑞/Δ𝜀vol when Δ𝑝′ = 0 (𝐶𝑀𝑆 and 𝐶𝑀𝑆𝐸) (6)

For undrained stress probes, 𝐺sec can be directly computed using Equation 3 since the
specimen experiences no volume changes under these conditions (Lings et al. 2000). According
to the wave propagation theory (Shirley and Hampton 1978), the small-strain elastic shear
modulus, 𝐺0 , and 𝑉s in the vertical direction are defined as:

𝐺0 = 𝜌𝑉s 2 ; 𝑉s = 𝑑/𝑡BE (7)

where 𝜌, 𝑑, and 𝑡BE are the soil mass density, wave travel distance, and wave travel time,
respectively. Herein, 𝑑 was determined from the distance between BE tips, and 𝑡BE was
estimated via the peak-to-peak method in the time domain.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 482

RESULTS

Response during 𝑲𝟎 -Reconsolidation

The quality of the soil specimens was evaluated via 𝜀vol and normalized change in void ratio,
Δ𝑒/𝑒0 , responses during recompression to 𝜎′v0 as suggested by Andresen and Kolstad (1979)
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

and Lunne et al. (2006), respectively. 𝜀vol varied from 0.20% to 0.63% and Δ𝑒/𝑒0 from 0.0055
to 0.0130, classifying the specimens as “very good to excellent” quality. These criteria to
evaluate soil specimen quality were employed in this study since HG sands are structured soils.
The synchronous measurements of 𝑉s during reconsolidation by reproducing the stress
history of the material provided data to define the evolution of 𝐺0 as shown in Figure 2. Herein,
the response is expressed as a function of the void ratio, 𝑒, and 𝑝′ and normalized with the
atmospheric pressure (𝑃a =101.3 kPa) and the void ratio function suggested by Shibata and
Soelarno (1978), yielding the following expression with a coefficient of determination, 𝑅 2 , of
0.93:
𝑛
𝐺0 𝑒 𝑝′
= 𝐴 (0.67 − )( ) (8)
𝑃a 1 + 𝑒 𝑃a

where 𝐴 and 𝑛 are material constants. The response of 𝐺0 for the 𝐾0 -loading path (○ symbol)
was smaller when compared with the 𝐾0 -unloading path (+ symbol) as indicated by the different
𝐴-values. This observation indicates that the direction of loading influences 𝐺0 , emphasizing the
importance of adequately modeling the material loading history during laboratory
reconsolidation.

Figure 2. Normalized 𝑮𝟎 response during reconsolidation by reproducing the stress history


of the HG sands.

Small-Strain Behavior

Results of the stress-probe tests are presented in Figure 3 in terms of stress-strain response in
the spaces Δ𝑞-𝜀sh , Δ𝑝′-𝜀vol , Δ𝑝′-𝜀sh , and Δ𝑞-𝜀vol up to the boundary of small strains (≤0.1%). In
the figure, magnified responses are shown in the insets, and each color represents paths 180°

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 483

apart in terms of total stress. In general, the stress-strain responses were non-linear and varied
with the direction of the applied stress. At each space of analysis, the initial response was stiffer
for the cases where Δ𝑞 or Δ𝑝′ positively increased in contrast to those where Δ𝑞 or Δ𝑝′
negatively increased. This increment in stiffness was more evident for the 𝑈𝑇𝑋𝐶 and 𝐶𝑀𝑆 paths
in the space Δ𝑞-𝜀sh than in the remaining paths and spaces of analysis. Note also that the onset of
stress-strain responses for 𝑈𝑇𝑋𝐶 and 𝐶𝑀𝑆 paths was similar. The stress-strain responses in
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figures 3(c) and 3(d) showed coupling at these strain levels between Δ𝑝′and 𝜀sh and between Δ𝑞
and 𝜀vol , respectively. More coupling was observed for the Δ𝑞-𝜀vol responses.

Figure 3. Small-strain behavior of HG sands in the spaces: (a) 𝚫𝒒-𝜺𝐬𝐡 , (b) 𝚫𝒑′-𝜺𝐯𝐨𝐥 , (c) 𝚫𝒑′-
𝜺𝐬𝐡 , and (d) 𝚫𝒒-𝚫𝜺𝐯𝐨𝐥 .

Figure 4 presents stiffness degradation responses of the shear, bulk, and cross-coupling
secant moduli. In the figure, the insets illustrate the stress path rotation angle for each stress
probe. The range of 𝐺0 , computed via BE testing at the end of the creep stage for the 𝑈𝑇𝑋𝐶,
𝑈𝑅𝑇𝑋𝐸, 𝐶𝑀𝑆, and 𝐶𝑀𝑆𝐸 paths, is included in Figure 4(a). In general, the stiffness responses
showed stress path and strain level dependency, being more evident for 𝐺sec as it was also
observed in the stress-strain responses in Figure 3. Note in Figure 4(a) that the onset of stiffness
degradation was strongly affected by the direction of loading. Stiffer responses were observed
for the 𝑈𝑇𝑋𝐶 (𝜃=130°) and 𝐶𝑀𝑆 (𝜃=135°) paths when compared with the 𝑈𝑅𝑇𝑋𝐸 (𝜃=35°) and
𝐶𝑀𝑆𝐸 (𝜃=45°) paths. The closeness in 𝜃-values for the 𝑈𝑇𝑋𝐶 and 𝐶𝑀𝑆 paths concords with the

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 484

measured responses. At strain levels larger than 1%, the effects of 𝜃 on 𝐺sec were not observed.
For the case of 𝐾sec , the effects of 𝜃 were only noticeable at the onset of stiffness degradation
and at large strains (>0.1%). This behavior can be attributed to the unusual abrupt displacements
that the specimen experienced under 𝐶𝑄𝐿 conditions at small strains. Note in Figure 3(b) the
stepped response for the 𝐶𝑄𝐿 path at strains less than 0.1%. The coupled secant moduli, 𝐽s sec
and 𝐽v sec , showed different degradation patterns, confirming that the influence of Δ𝑞 on 𝜀vol is
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

larger than that of Δ𝑝′ on 𝜀sh . Uncoupling behavior was observed at strains of approximately
0.2%, 0.2%, and 0.03% for the 𝐶𝑄𝐿, 𝐶𝑀𝑆, and 𝐶𝑀𝑆𝐸 paths, respectively. The stiffness
degradation of 𝐽v sec clearly showed stress path rotation angle effects. A stiffer response was
obtained for the 𝐶𝑀𝑆 (𝜃=135°) path when compared with the 𝐶𝑀𝑆𝐸 (𝜃=45°) path. The effects
of 𝜃 on 𝐽s sec were similar to those observed for 𝐾sec .

Figure 4. Stiffness degradation responses of (a) shear, (b) bulk, and (c) cross-coupling
secant moduli of HG sands.

SUMMARY AND CONCLUSIONS

This paper presented the results and analyses of triaxial stress probes with internal
instrumentation, BE and HE displacement transducers, conducted to examine the small-strain
behavior and effects of stress path rotation angle of HG sands in Central Florida. Responses
during 𝐾0 -reconsolidation were employed to evaluate the quality of tested specimens and define
the evolution of 𝐺0 via BE tests. The small-strain behavior of the material was studied under
undrained and drained triaxial conditions following compression and extension stress paths and
evaluated in terms of stress-strain response with emphasis on stiffness degradation of the shear,
bulk, and cross-coupling moduli. The effects of stress path rotation angle were considered when
evaluating the small-strain stiffness responses. From the information presented herein, the
following conclusions can be drawn:
1. The response of the small-strain elastic shear modulus under 𝐾0 -reconsolidation paths of
HG sands is influenced by the direction of loading as indicated by the different 𝐴-values
obtained for the 𝐾0 -loading and unloading paths evaluated herein, emphasizing the
importance of adequately modeling the material loading history during laboratory
reconsolidation.
2. The stress-strain and stiffness degradation responses obtained via undrained and drained
stress probes show that the small-strain behavior of HG sands is non-linear, strain level

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 485

and stress path dependent. The effects of the stress path rotation angle on the small-strain
stiffness behavior are evidenced at the onset of responses and decrease with stress level,
being more pronounced for the shear secant moduli. Stiffer responses were observed for
𝜃-values approaching 180°.
3. The evaluation of coupling behavior of HG sands shows that the influence of the deviator
stress on the volumetric strain is larger than that of the effective mean normal stress on
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the shear strain as indicated by the different degradation patterns of the coupled secant
moduli, 𝐽s sec and 𝐽v sec . Uncoupling behavior was observed at strains of approximately
0.2%, 0.2%, and 0.03% for the 𝐶𝑄𝐿, 𝐶𝑀𝑆, and 𝐶𝑀𝑆𝐸 paths, respectively.

ACKNOWLEDGEMENTS

Financial support was provided by the Colombian Administrative Department of Science,


Technology and Innovation (COLCIENCIAS), Scholarship Program No. 757-2016, and the
Florida Department of Transportation (FDOT), Grant No. BDV24 TWO 977-29. The support of
the funding agencies is greatly appreciated.

REFERENCES

Andresen, A., and Kolstad, P. (1979). “The NGI 54 mm Sampler for Undisturbed Sampling of
Clays and Representative Sampling of Coarser Materials.” Int. Symp. Soil Sampl., Singapore,
13–21.
Aparicio-Ortube, A. J., Arboleda-Monsalve, L. G., and Zapata-Medina, D. G. (2022). “Intralayer
Variability and Compressibility of Hawthorn Group Soils.” Geotech. Geol. Eng.
Aparicio-Ortube, A. J., Arboleda-Monsalve, L. G., Zapata-Medina, D. G., and Jones, L. (2021).
“Stress History Effects on Shear Stiffness Degradation under Compression Paths of
Hawthorn Group Clays in Central Florida.” IFCEE 2021, American Society of Civil
Engineers, 224–233.
Atkinson, J. H., Richardson, D., and Stallebrass, S. E. (1990). “Effect of Recent Stress History
on the Stiffness of Overconsolidated Soil.” Géotechnique, 40(4), 531–540.
Becker, D. E., Crooks, J. H. A., Been, K., and Jefferies, M. G. (1987). “Work as a Criterion for
Determining In Situ and Yield Stresses in Clays.” Can. Geotech. J., 24(4), 549–564.
Bjerrum, L. (1967). “Engineering Geology of Norwegian Normally-Consolidated Marine Clays
as Related to Settlements of Buildings.” Géotechnique, 17(2), 83–118.
Burland, J. B. (1989). “Ninth Laurits Bjerrum Memorial Lecture: ‘Small is Beautiful’—The
Stiffness of Soils at Small Strains.” Can. Geotech. J., 26(4), 499–516.
Cho, W., and Finno, R. J. (2010). “Stress-Strain Responses of Block Samples of Compressible
Chicago Glacial Clays.” J. Geotech. Geoenvironmental Eng., 136(1), 178–188.
Cho, W., Holman, T., Jung, Y., and Finno, R. (2007). “Effects of Swelling During Saturation in
Triaxial Tests in Clays.” Geotech. Test. J., 30(5), 378–386.
Finno, R. J., and Cho, W. (2011). “Recent Stress-History Effects on Compressible Chicago
Glacial Clays.” J. Geotech. Geoenvironmental Eng., 137(3), 197–207.
Finno, R. J., and Kim, T. (2012). “Effects of Stress Path Rotation Angle on Small Strain
Responses.” J. Geotech. Geoenvironmental Eng., 138(4), 526–534.
Green, R. C., Williams, C. P., Bambach, P. W., Hannon, L. M., Apolinar, B., Campbell, K. M.,
and Dyer, S. B. (2015). Text to Accompany Geologic Map of the USGS Orlando 30 x 60
Minute Quadrangle, Central Florida.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 486

Høeg, K., Dyvik, R., and Sandbækken, G. (2000). “Strength of Undisturbed versus Reconstituted
Silt and Silty Sand Specimens.” J. Geotech. Geoenvironmental Eng., 126(7), 606–617.
Huang, A. B., and Huang, Y. T. (2007). “Undisturbed Sampling and Laboratory Shearing Tests
on a Sand with Various Fines Contents.” Soils Found., 47(4), 771–781.
Kenney, T. C. (1964). “Sea-Level Movements and the Geologic Histories of the Post-Glacial
Marine Soils at Boston, Nicolet, Ottawa and Oslo.” Géotechnique, 14(3), 203–230.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Kim, T., Zapata-Medina, D. G., and Vega-Posada, C. A. (2015). “Analysis of Bender Element
Signals during Triaxial Testing.” Rev. Fac. Ing. Univ. Antioquia, (76), 107–113.
Kim, Y., Nam, B. H., Shamet, R., Soliman, M., and Youn, H. (2020). “Development of Sinkhole
Susceptibility Map of East Central Florida.” Nat. Hazards Rev., 21(4), 04020035.
Ladd, C., and Lambe, T. (1964). “The Strength of ‘Undisturbed’ Clay Determined from
Undrained Tests.” Lab. Shear Test. Soils, ASTM International, 342–371.
Lings, M. L., Pennington, D. S., and Nash, D. F. T. (2000). “Anisotropic Stiffness Parameters
and their Measurement in a stiff Natural Clay.” Géotechnique, 50(2), 109–125.
Lunne, T., Berre, T., Andersen, K. H., Strandvik, S., and Sjursen, M. (2006). “Effects of Sample
Disturbance and Consolidation Procedures on Measured Shear Strength of Soft Marine
Norwegian Clays.” Can. Geotech. J., 43(7), 726–750.
Nam, B. H., and Shamet, R. (2020). “A Preliminary Sinkhole Raveling Chart.” Eng. Geol., 268,
105513.
Santagata, M., Germaine, J. T., and Ladd, C. C. (2005). “Factors Affecting the Initial Stiffness of
Cohesive Soils.” J. Geotech. Geoenvironmental Eng., 131(4), 430–441.
Scholey, G., Frost, J., Lo Presti, D., and Jamiolkowski, M. (1995). “A Review of Instrumentation
for Measuring Small Strains During Triaxial Testing of Soil Specimens.” Geotech. Test. J.,
18(2), 137.
Scott, T. M. (1988). The Lithostratigraphy of the Hawthorn Group (Miocene) of Florida.
Tallahassee, Florida. Florida Geological Survey.
Scott, T. M. (2001). Text to Accompany the Geologic Map of Florida. Tallahassee, Florida.
Florida Geological Survey.
Scott, T. M., MacGill, P. L., Eddy, W. H., Davis, B. E., and Sullivan, G. V. (1981). The
Hawthorn Formation of Central Florida. Tallahassee, Florida. Florida Geological Survey.
Shamet, R., Nam, B. H., and Horhota, D. (2018). “Development of a Point-Based Index for
Sinkhole Vulnerability Evaluation in Central Florida’s Karst Terrain.” IFCEE 2018,
American Society of Civil Engineers, 212–221.
Shibata, T., and Soelarno, D. S. (1978). “Stress-Strain Characteristics of Clays under Cyclic
Loading.” Proc. Japan Soc. Civ. Eng., 1978(276), 101–110.
Shirley, D. J., and Hampton, L. D. (1978). “Shear‐wave Measurements in Laboratory
Sediments.” J. Acoust. Soc. Am., 63(2), 607–613.
Soliman, M. H., Arboleda-Monsalve, L. G., and Nam, B. H. (2020). “Effects of Intergranular
Strains of Hypoplasticity Models on Sinkhole-Induced Ground Deformations.” Geo-
Congress 2020, American Society of Civil Engineers, 265–274.
Upchurch, S., Scott, T. M., Alfieri, M. C., Fratesi, B., and Dobecki, T. L. (2019). The Karst
Systems of Florida. Cave and Karst Systems of the World, Springer International Publishing,
Cham.
Vaid, Y. P., Sivathayalan, S., and Stedman, D. (1999). “Influence of Specimen-Reconstituting
Method on the Undrained Response of Sand.” Geotech. Test. J., 22(3), 187–195.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 487

Weary, D. J. (2015). “The Cost of Karst Subsidence and Sinkhole Collapse in the United States
Compared with Other Natural Hazards.” 14th Sink. Conf., University of South Florida, 433–
446.
Yamamuro, J. A., and Wood, F. M. (2004). “Effect of Depositional Method on the Undrained
Behavior and Microstructure of Sand with Silt.” Soil Dyn. Earthq. Eng., 24(9–10), 751–760.
Zapata-Medina, D. G., Finno, R. J., and Vega-Posada, C. A. (2014). “Stress History and
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Sampling Disturbance Effects on Monotonic and Cyclic Responses of Overconsolidated


Bootlegger Cove Clays.” Can. Geotech. J., 51(6), 599–609.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 488

Monotonic Behavior of Ledge Point Calcareous Sands with Increasing Particle Crushing

Wenjing Cai, Ph.D.1; and Cassandra J. Rutherford, Ph.D., P.E., M.ASCE2


1
Assistant Professor, College of Environment and Civil Engineering, Chengdu Univ. of
Technology, Chengdu, China. Email: caiwenjing@cdut.edu.cn
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Assistant Professor, Dept. of Civil, Construction, and Environmental Engineering, Iowa State
Univ., Ames, IA. Email: cassier@iastate.edu

ABSTRACT

Calcareous sands have different engineering properties than clays and silica sands due to
their high particle angularity and in situ void ratios. The characteristics and stress-strain response
of calcareous sands are necessary for the design and long-term maintenance of offshore
infrastructure, such as piers, bridges, and offshore renewable energy systems. The monotonic
stress-strain response of calcareous sands from Ledge Point Australia is examined in this study,
as well as the effect of increasing particle damage and crushing severity on loading behavior.
The calcareous specimens were crushed by both consolidation and shearing methods to obtain
varying amounts of crushed material. Results indicated that as the content of particle breakage in
the sample matrix increased, the crushed sands contracted less and exhibited a stronger dilative
behavior during shearing. However, the highest shear strength was obtained by the sample with
relative particle breakage (Br) of 0.05, rather than the sample with the highest crushing
content.

INTRODUCTION

Calcareous sediments mostly originate from biochemical processes attributed to coral reef
formation, sedimentation of skeletal debris, and chemical precipitation of particles (Murff, 1987;
Milliman et al., 1974). This type of sand shows high compressibility during loading, up to 100
times higher than silica sand (Nauroy & LeTirant, 1982), attributed to particle breakage from low
hardness and intraparticle voids within the grains. During driving, calcareous sands can
experience dramatic particle damage, resulting in excessive foundation settlement. It has been
reported that measured load capacities of offshore construction within calcareous sediments were
approximately 20% lower than predicted capacities. This underestimation of load capacity often
occurred in sands with high potential for particle damage. Particle breakage is understood to
influence stress-strain response and other geotechnical behavior. Hence, it is necessary to
investigate the effect of particle breakage of calcareous sands and the resulting changes to its
engineering behaviors. This research is focusing on the effect of particle breakage on static
stress-strain response. Calcareous sands from Ledge Point, Australia are the focus of this study.
Triaxial undrained tests were initially performed on the uncrushed calcareous sands to
investigate stress-strain behavior. Two methods were used to evaluate the crushability of
calcareous sands, and to initiate the different particle crushing contents. A series of
reconstituted calcareous specimens were prepared with different crushing content in order to
investigate the effect of increasing crushing content on the static stress-strain response for
calcareous sands.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 489

MATERIALS

Australia is surrounded by the Indian and Pacific oceans, and Western Australia (WA) is the
largest state in Australia, occupying the entire western third of the country. Calcareous sands
selected for the present study are from Ledge Point, a small coastal township 105 km north of
Perth, Western Australia. Sediments from Ledge Point are primarily coastal aeolian deposited
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

calcareous soils. As a part of the Quindalup dune system, calcareous sand dunes were formed
along the shoreline of the Perth basin 10,000 years ago (Sharma, 2004). Different water depth
leads to different grain sizes and compositions in skeletal grain types, which potentially affect
their engineering properties (Beemer & Levrec, 2018; Beemer et al., 2019; Beemer & Sadekov,
2019). In this study, calcareous sands were collected from the seabed with a 10-meter water
depth.
According to the particle size distribution (PSD) curve of Ledge Point (LP) sands (Figure 1),
the grain size which was distributed from 2 mm (No. 10 sieve) to 0.075 mm shows a more
uniform gradation with the Cu and Cc values of 2.40 and 1.20, respectively. The LP sands had
more than 50% of coarse fraction passing No. 4 sieve and had less than 5% passing No. 200, that
classified as poorly graded sands (SP). Specific gravity of LP sand is 2.76, and calcium
carbonate content is 93%.

100
Cummulative Percentage of Passing, %

80

60

40

20
LP Sand
0
0.01 0.10 1.00 10.00
Sieve Opening (mm)

Figure 1. Particle size distribution of calcareous sands from Ledge Point, Australia.

METHODOLOGY

Particle crushing

In this study, the calcareous sands are crushed by two methods: consolidated drained (CD)
triaxial tests and one-dimensional (1-D) consolidation tests. It was shown by Shahanazari and
Rezvani (2013) that the drained conditions led to more particle breakage for crushable sands than
that of undrained conditions; therefore, the triaxial tests to obtain crushed samples were
conducted in a drained condition

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 490

The testing specimen for triaxial CD test was a cylindrical shape with a diameter of 7.11 cm
and a height of 14.22 cm. A thicker than typical latex membrane (0.06 cm thickness) was used to
prevent membrane breakage or damage due to the high angularity of calcareous sand. Confining
pressure levels of 100 kPa, 300 kPa, and 400 kPa, were applied to obtain different particle
crushing contents.
For 1-D consolidation test, the sample size was 6.35 cm in diameter and 2.54 cm in height,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

and the stress levels selected for crushing were 2400 kPa and 4800 kPa. For both crushing
methods, calcareous sand samples were prepared by dry pluviation method to obtain the desired
relative density (Dr = 50%). The samples were prepared and tested in a dry condition to
maximize the particle crushing.
Particle size distributions of uncrushed LP sand and the crushed sands obtained from 1-D
consolidation (ODC) and triaxial CD (TxCD) tests are presented in Figure 2. Relative particle
𝐵
breakage (Br) for each test was determined based on Hardin’s method (1985): 𝐵𝑟 = 𝑡 , where Br
𝐵𝑝
represents the relative particle breakage, Bt is the area between the initial and final gradation
curves, and Bp is the area between the original particle size distribution curve and vertical line for
particle size of 0.075 mm (U.S. sieve No. 200).
Hardin (1985) illustrated that the breakage of particles terminated when the soil gradation
curve achieved a stable condition. He also came up with a method to assess particle crushing.
This method is still used in recent investigations. According to Hardin’s method, the changes in
the particle size distribution curve before and after crushing are determined and the final stable
condition was reached when all particles were smaller than 0.075 mm. The results for relative
particle breakage at the two stress conditions are summarized in Table 1.

Table 1. Relative particle breakage (Br) of Ledge Point sand for 1-D consolidation tests and
triaxial CD tests.

Test ODC TxCD


Stress (kPa) 2,400 4,800 100 300 400
Br 0.05 0.07 0.02 0.08 0.14

Figure 2. (a) particle size distribution of virgin and consolidation crushed Ledge Point
sands; (b) particle size distribution of virgin and shear crushed Ledge Point sands.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 491

Figure 2 shows PSD curves for LP sands prior to and after crushing. These plots indicate that
compression and shear loading caused particle breakage at different particle size ranges. The
gradation change caused by compression loading occurred at two particle size ranges: 0.2 mm to
0.6 mm and 0.075 mm to 0.17 mm whereas the gradation change caused by shear loading was
observed between the particle size of 0.6 mm and 0.075 mm. Also observed from the crushing
tests was that compression loading caused less particle breakage with large loading stresses,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

while more crushing occurred by shear loading with relatively lower loading stresses.

Monotonic shear test

To investigate the effect of increasing particle crushing on the shear strength and stress-strain
response, triaxial consolidated undrained (CU) tests were conducted on the reconstituted samples
with 400 kPa confining pressure using different particle breakage content. Previous research
shows that calcareous sands are sustainable to crushing under undrained shearing (Hardin, 1985;
Hyodo et al., 1999). However, this research focuses on the effect of increase particle content
after initial crushing. A series of the LP sand test specimens were prepared with increasing
particle breakage contents obtained previously using consolidate and drained triaxial testing (i.e.
Br = 0.02, 0.05, 0.07, 0.08, and 0.14). Specimens were prepared by dry pluviation method at 50%
relative density. The dry sand sample was 7.11 cm in diameter and 14.22 cm in height. The
specimen was flushed with CO2 gas followed by flushing de-aired water to achieve above 95%
saturation. Flushing CO2 gas speeds up the saturation process by exhausting the air inside the
sample and dissolving in the injected de-aired water. The sample was consolidated at 200 kPa
isotropic pressure for 15 min, and then sheared at a confining pressure of 400 kPa until 15%
axial strain. The shearing rate was 1% strain per hour.

TEST RESULTS

As shown in the stress-strain curves (Figure 3.a), the uncrushed (Br = 0) LP sand resulted in
the lowest shear strength within 15% axial strain. As the particle breakage content increased
from 0 to 0.02, there was a clear increase of shear strength change which continued to increase
until Br =0.05. Beyond Br =0.05 the increase of crushed material indicated a clear decrease in
shear strength. Except for the samples of Br = 0 and 0.08, other specimens exhibited brittle
behavior by achieving the maximum strength at 11% to 14% strain, followed by slightly strain
softening beyond maximum strength. As shown by the excess pore pressure generation vs strain
plots (Figure 3.b), the uncrushed sample had the highest pore pressure generation with
decreasing pore pressure generation with increasing crushed material.
In the case of the LP sand, the shear strength increased when particle breakage increased for
the initial breakage content (Br = 0.02 and 0.05). However, shear strength decreased after Br
=0.05. As illustrated in Figure 4, the strength of the crushed material even at Br = 0.14 was
higher than the strength of uncrushed sand (Figure 4). This indicates that the crushed particles
effect the strength response of the sands with an initial increase; however, with continued
crushing, the sand shows a decrease in strength.
Figure 5 shows plots of the uniformity coefficient (Cu) for each crushed sand with its
corresponding peak strength. Samples achieved the highest strength (Br = 0.02 and 0.05)
obtained the similar Cu value around 2.49; however, other samples below and beyond than this
Cu value were all gained a relative low strength. Therefore, there is an optimum Cu range for
crushed LP sand to achieve a higher strength. The particle size distribution of crushed calcareous

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 492

sand plays a role in the development of strength and should be considered when analyzing the
effect on shear strength with increasing particle breakage, especially for the crushable material
with fine gradation.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. LP Sand with increasing particle breakage (a) stress-strain response; (b) excess
pore water pressure.

2500
Maximum Shear Strength (kPa)

2000 LP Sand

1500

1000

500

0
0 0.05 0.1 0.15
Br

Figure 4. Shear strength with increasing Br values for LP sand.

Figure 5. Correlation between uniformity coefficient (Cu) and shear strength for LP sand.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 493

CONCLUSION

Triaxial CU tests were performed on calcareous sand samples with a confining pressure of
400 kPa to investigate the static loading behavior with the increasing particle breakage. To
obtain different particle crushing content, 1-D consolidation test and triaxial CD test were used.
Based on the crushing results, it is found that shear loading is more effective for crushing LP
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

sand particles than 1-D compression.


For the stress-strain response, the measured shear strength did not present a linear
relationship with Br value. Higher shear strength of the LP sands was obtained by the sample
with Br = 0.05 and then shear strength decreased with higher crushing content. Therefore,
crushing content of calcareous sands appears to influence shear strength. Also, the effects of
particle gradation and particle arrangement should be considered. Results indicated that the
strength for LP sands may be related to the uniformity coefficient (Cu), an optimum range of
uniformity coefficient value (around Cu = 2.5) contributed to the highest measured shear
strength.

REFERENCES

Beemer, R. D., Bandini-Maeder, A., Shaw, J., and Cassidy, M. J. (2019). Volumetric Particle
Size Distribution and Variable Granular Density Soils. Geotechnical Testing Journal. 43(2),
17.
Beemer, R. D., Sadekov, A., Levrec, U., Shaw, J., Bandini-Maeder, A., and Cassidy, M. J.
(2019). Impact of Biology on Particle Crushing in Offshore Calcareous Sediments: Geo-
Congress 2019. 640-650.
Beemer, R. D., Bandini-Maeder, A. N., Shaw, J., Levrec, U., and Cassidy, M. J. (2018). The
Granular Structure of Two Marine Carbonate Sediments. OMAE2018, ASME, Madrid.
Hardin, B. O. (1985). Crushing of soil particles. Journal of Geotechnical Engineering. 111 (10),
1177–1192.
Hyodo, M., Aramaki, N., Nakatay, Y., Inoue, S., and Hyde, A. F. L. (1999). Particle Crushing
and Undrained Shear Behaviour of Sand: Cupertino, CA, ETATS-UNIS, Int. Soc. Offshore.
Polar. Eng.
Milliman, J. D., Müller, G., and Förstner, U. (1974). Carbonates and the ocean. In Recent
Sedimentary Carbonates (pp. 3-15). Springer, Berlin, Heidelberg.
Murff, J. D. (1987). “Pile capacity in calcareous sands: State of the art.” J. Geotech. Eng., ASCE,
113(5), 490–507.
Nauroy, J. F., and Le Tirant, P. (1982). Comportement des sédiments marins carbonates. Oil &
Gas Science and Technology – Rev. IFP 37(2): 149–155.
Sharma, S. S. (2004). Characterization of cyclic behavior of calcite cemented calcareous soils.
Ph.D. thesis, Univ. Western Australia, Perth, Australia.
Shahnazari, H., and Rezvani, R. (2013). Effective parameters for the particle breakage of
calcareous sands: An experimental study. Engineering Geology, 159, 98-105.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 494

Laboratory Evaluation of Small Strain Elastic Parameters of Coal Ash from Bender
Element Tests

C. S. S. U. Srikanth1; B. J. Ramaiah, Ph.D., A.M.ASCE2; and Murali Krishna, Ph.D., M.ASCE3


1
Research Scholar, Dept. of Civil and Environmental Engineering, Indian Institute of
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Technology Tirupati, Tirupati, India. ORCID: https://orcid.org/0000-0002-3168-800X.


Email: ce19d503@iittp.ac.in
2
Assistant Professor, Dept. of Civil and Environmental Engineering, Indian Institute of
Technology Tirupati, Tirupati, India (corresponding author).
ORCID: https://orcid.org/0000-0001-7480-4499. Email: janakiram@iittp.ac.in
3
Professor, Dept. of Civil and Environmental Engineering, Indian Institute of Technology
Tirupati, Tirupati, India. ORCID: https://orcid.org/0000-0002-5422-2501.
Email: amk@iittp.ac.in

ABSTRACT

Increased industrialization, urbanization, and economic growth of developing countries like


India have demanded more energy requirements. Currently, thermal power plants are the
dominating power generation sources. Accordingly, coal ash management has become a huge
challenge in India and globally. Mass utilization of coal ash is possible in geotechnical
applications and hence requires a proper understanding of its elastic parameters such as Young's
modulus, shear modulus, and bulk modulus. The present study investigated the effect of sample
preparation, initial void ratio, confining pressure, and S and P wave travel time measurement on
the elastic parameters of coal ash using bender element tests. Additionally, power models for the
maximum shear modulus (Gmax) and maximum Young’s modulus (Emax) were developed, and the
fitting constants, A and n, for the relationship were evaluated. These relationships are useful as
input parameters in static and seismic analysis of various geotechnical structures constructed
with coal ash.

INTRODUCTION

Increasing demands for electricity have led to a rapid increase in coal ash generation at
thermal power plants. This scenario is almost similar in both developed and developing
countries. The effective utilization of coal ash in various geotechnical constructions requires a
comprehensive understanding of its mechanical properties in the linear and non-linear range of
deformation. The maximum small strain shear modulus (Gmax) is a predominant input parameter
not only for the dynamic and liquefaction analysis of geotechnical structures but for most critical
state soil mechanics (CSSM) models. Evaluation of the Gmax and other elastic parameters
becomes more significant due to the susceptibility of coal ash to liquefaction (Jakka et al. 2010;
Zand et al. 2009). Field techniques to evaluate the stiffness of the geomaterial include Spectral
analysis of surface waves (SASW) and Multi-channel analysis of surface waves (MASW), which
allow the development of shear velocity (Vs) profiles. Various researchers reported the measured
field Vs profiles of the coal ash storage facilities through SASW and MASW techniques (Parhi et
al. 2020; Ramaiah et al. 2010).
Resonant column and bender element tests complement the field Vs testing with sophisticated
laboratory evaluation procedures. The bender element test is considered a simple, quick, and

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 495

effective substitute laboratory method to measure the Vs and subsequently to calculate the Gmax
of the soil (Janusz et al. 2021; Moldovan et al. 2016). There is abundant literature on the bender
element testing on typical soils of sands and clays, but very few studies have been conducted on
coal ash (Bachus et al. 2019; Choo et al. 2016). The present study presents the experimental data
on the small strain elastic parameters of the coal ash using the bender extender elements, which
can transmit and receive both S-waves and P-waves (Lings and Greening 2001). This study
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

investigated the effect of the method of measurement of travel time of S and P-wave (Ts and Tp),
void ratio (e), effective confining pressure (σ’), and saturation on the elastic parameters of the
coal ash. In addition, power models describing the relationship between the elastic parameters
and the effective confining pressure and void ratio were developed for coal ash, Poison’s ratio,
bulk modulus, and the fitting constants were evaluated.

CHARACTERIZATION OF COAL ASH

The coal ash used in the present study was collected from an ash pond in National Thermal
Power Corporation (NTPC), Visakhapatnam, India. Table 1 presents the results of physical
characterization of the collected coal ash sample, including specific gravity, particle size
distribution, compaction characteristics (based on modified Proctor test), and loss on ignition
tests. The liquid limit (LL) of the coal ash sample has been determined using the fall cone
method as per ASTM D4318-17 and the ko stress method (Sridharan et al. 2000). It is observed
that the LL obtained from these two methods is almost similar value. Based on the physical
properties, the collected coal ash material is classified as sandy silt type with low plasticity as per
the unified soil classification system prescribed in ASTM D2487-17.

Table 1. Geotechnical characterization of the coal ash considered in the present study

Property Value
Specific gravity 2.1
Sand (%) 38
Silt (%) 59
Clay (%) 3
Coefficient of uniformity, Cu 10.5
Coefficient of curvature, Cc 0.6
Classification ML (Sandy Silt)
Liquid limit, LL (%) 27.2
Optimum moisture content, OMC (%) 18
Maximum dry density, MDD (g/cc) 1.32
Loss on ignition (%) 2.91

TESTING EQUIPMENT

The bender element tests were conducted using a modified triaxial cell base and top cap
connected to the bender scope, consisting of an inbuilt function generator and a digital
oscilloscope. The two piezo-metric ceramic elements are provided in capsules made with
titanium to reduce their weight with a protrusion height of 1.5 mm. A typical test setup is shown
in Figure 1. The bender element in the top cap acts as a transmitter to produce a shear wave,

© ASCE @seismicisolation
@seismicisolation
Bender elements connected to top cap and base pedestal

Geo-Congress 2023 GSP 340 496

while the bender element connected to the base pedestal acts as a receiver. The bottom element
transmits a mechanical excitation producing compressional wave which is captured and
converted to an electrical impulse by the top bender element. Pressure volume controller to regulate
Bender elements connected to top cap and base pedestal

Bender elements connected to top cap and base pedestal


cell and back pressure
Bender elements connected to top cap and base pedestal
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
1 2 Pressure volume controller to regulate
cell and back pressure 4
2
Pressure volume controller to regulate
cell and back pressure
3
Bender Scope
Benderwith inbuilt
Scope with inbuilt function generator
function generator Pressure
and andvolume
oscilloscope oscilloscope
controller
Data acquisition to Data
using PC acquisition using PC
regulate
cell and back pressure
1 Bender elements connected to top cap
Bender Scope with inbuilt function generator and oscilloscope Data acquisition using PC
and base pedestal
2 Pressure volume controller for
3 4 regulating confining pressure and
back
Bender Scope with inbuilt function generator pressure Data acquisition using PC
and oscilloscope
3 Bender scope with inbuilt function
generator and oscilloscope
4 Data acquisition using computer

Figure 1. Schematic diagram of the bender element test setup used for the present study

METHODOLOGY

The bender element testing was conducted as per ASTM D8295-19. The calculated sample
weight was separated into five equal parts, and water was added to each layer per the required
water content. The coal ash-water mixture was allowed to equilibrate for 24 hours. The specimen
is prepared in a two-part split mold in five layers. Once the specimen was prepared, it was
extracted from the mold and carefully placed on the base with a bender element. Using a rubber
membrane extender, the sample was isolated from its surroundings. Finally, the top porous stone
and top cap with the bender element were placed carefully in line with the bottom bender
element to avoid the inversion problem of the received wave. To ensure a better connection
between the ceramic and the coal ash, a protrusion the size of the piezo element was prepared on
the sample before adding the top cap (ASTM D8295-19).
Techniques for pre-saturation, such as CO2 saturation and vacuum saturation, were used. The
specimen was saturated using the incremental backpressure technique. The specimen was
considered saturated when B-value was greater than 0.95. The sample was then consolidated at
different effective confining pressures. After the consolidation, the signals were sent using a
function generator in the bender scope. The same specimen was used for the bender element
testing at various confining pressures in order to maintain consistency between test results and
for comparative purposes. However, before proceeding to the next confining pressure stage, the
B-check was performed to ensure the sample was saturated.
In the case of the dry tamping technique, the sample was prepared on the base pedestal. After
the sample was prepared, a negative suction of not more than 35 kPa was applied to the sample
for at least two minutes via a back-pressure line to help the specimen retain stability on its own.
After applying confining pressure, sufficient time was allowed for the sample to stabilize, and
the signals were shot at different frequencies ranging from 1 kHz to 50 kHz (Gu et al. 2013). Due
to its simplicity in measuring travel time, the sinusoidal waveform was used as the input signal

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 497

(Ogino et al. 2015), and multiple signals were captured at varied input frequencies at each stage
with a capture duration of 5 ms. Stacking of signals and a strong cutoff low pass filter was
employed to get a better receiver signal output, which was measured with an accuracy of ± 1
mV. The testing scheme employed for the present study is tabularized in Table 2.

Table 2. Test matrix of the current study


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Void ratio, Relative State of Method of Confining Input


e compaction, coal ash sample pressure (kPa) frequency
RC (%) preparation (kHz)
0.683 95 Saturated Moist tamping 50, 100, 200 1, 2, 3, 5, 10,
20, 50
0.665 96 Dry and Dry tamping 10, 50, 75, 100, 2, 3, 5, 10, 20,
Saturated 200, 400, 600 50
0.926 83 Dry Dry tamping 0, 50, 100, 200, 2, 10
500, 800
1.284 70 Dry Dry tamping 0, 20 2, 5, 10, 50

RESULTS AND DISCUSSION

When the excitation frequency is lower than the natural frequency of the material, a small dip
in the received signal, commonly referred to as the near field effect, is detected, complicating the
determination of travel time. In the case of s-wave transmission, this near field effect is one of
the most challenging aspects of signal interpretation, processing, and analysis of gathered data.
The interference of the p-wave part of the signal causes ambiguity in the s-wave travel time
measurement. A parameter known as wave path (distance between the source and receiver
bender elements, Ltt) to wavelength ratio (Ltt/λ) is calculated as the product of transmission
frequency (ft) and travel time (∆𝑡) to evaluate the near field effect (Gu et al. 2013) .

𝐿𝑡𝑡
= 𝑓𝑡 ∗ ∆𝑡 (1)
λ

It is reported that an Ltt/λ value of greater than 2 produces a clear signal free of near-field
effect (Gu et al. 2013). Figure 2 presents the output received s-wave excitation signals for a
prepared at 95 % of MDD in saturated and dry states at different frequencies and corresponding
Ltt/λ ratios. It can be noted that for an Ltt/λ > 2, the output signal becomes more coherent.
However, if the Ltt/λ exceeds a certain limiting value, the signal gets distorted and diminished
due to high damping when it reaches the receiver element. In the present study, the general range
of Ltt/λ value for coal ash was observed as 2 <Ltt/λ< 7.
The problem is negligible in the case of p-wave measurement, as compressional waves are
generally free of near-field effects. However, after saturation, because the compressional wave
travels faster in water (at a velocity of 1500 m/s), low-frequency components diminish, as shown
in Figure 3 (b), and the receiver element captures the waves that travel through the water. As a
result, determining the p-wave velocity of the material is impossible after the saturation.
However, because the B-check technique is just a laboratory process that cannot be transferred to
the field to explore saturation, saturated p-wave velocity measurement can be used as a trigger in
the field to determine the saturation of the soil.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 498

Effect of travel time measurement technique

In the present study, the peak-to-peak method (p-p), start-to-start method (s-s), inversion
method (inv), and cross-correlation method were adopted to find the travel time between source
and receiver signals (Murillo et al. 2011). Figure 4 is a trial representation of the time domain
analysis conducted in the study. As illustrated in Figure 4, the travel time in the start-to-start
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

approach was calculated as the distance between the beginning of the source signal and the
position zero of the received signal after inversion. The peak-to-peak method considered the
travel time between the first peaks of source and receiver signals, respectively. The inversion
method considered the distance between the source signal's start time and the start point of the
receiver signal immediately after inversion. In general, the start-to-start method is constant
across input frequencies, but the peak-to-peak method suffers from inconsistency between input
frequencies despite being closer to objectivity. A mid-range value that closely resembles the field
scenario is produced by the inversion method.
The time-domain signals are converted into the frequency domain and multiplied to obtain a
single dominant waveform in the cross-correlation method. The frequency-domain signal is
converted to a time-domain signal by inverse Fourier transform after correlation, as shown in the
equations below, and the travel time is measured as the time for the first peak (CC-1) and
maximum peak (CC-2). In some cases, the first and maximum peaks coincided to form a single
largest peak (CC).
(a) (b)
ft = 50 kHz, Ltt/l = 19 ft = 50 kHz, Ltt/l = NA

ft = 30 kHz, Ltt/l = 12 ft = 20 kHz, Ltt/l = 7.5


Signal Output

ft = 10 kHz, Ltt/l = 4.5


ft = 20 kHz, Ltt/l = 8.2
Signal Output

ft = 7 kHz, Ltt/l = 3.4

ft = 10 kHz, Ltt/l = 4.2


ft = 6 kHz, Ltt/l = 3

ft = 8 kHz, Ltt/l = 3.4


ft = 5 kHz, Ltt/l = 2.1
ft = 7 kHz, Ltt/l = 3
ft = 5 kHz, Ltt/l = 2.3 ft = 3 kHz, Ltt/l = 1.5
ft = 3 kHz, Ltt/l = 1.3 ft = 2 kHz, Ltt/l = 1
ft = 2 kHz, Ltt/l = 0.8
ft = 1 kHz, Ltt/l = 0.5
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Time (ms) Time (ms)

Figure 2. Received s-wave signals in (a) dry coal ash and (b) saturated coal ash.
(a) (b)
ft = 50 kHz

ft = 50 kHz

ft = 30 kHz
Signal Output

Signal Output

ft = 30 kHz
ft = 20 kHz
ft = 25 kHz
ft = 10 kHz ft = 20 kHz
ft = 8 kHz ft = 15 kHz
ft = 7 kHz ft = 10 kHz
ft = 8 kHz
ft = 5 kHz ft = 6 kHz
ft = 5 kHz
ft = 3 kHz ft = 3 kHz

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Time (ms) Time (ms)

Figure 3. Received p-wave signals in (a) dry coal ash and (b) saturated coal ash.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 499

The basic conceptualization of the CC method is described in Equations (2) to (4) given
below:

𝐿(𝑓) = 𝐹𝐹𝑇(𝑋(𝑡)); 𝑀(𝑓) = 𝐹𝐹𝑇(𝑌(𝑡)) (2)

𝐶𝐶𝑥𝑦 = 𝑀∗ (𝑓) ∗ 𝐿(𝑓) (3)


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

𝐼𝐹𝐹𝑇(𝐶𝐶𝑥𝑦 ) = 𝐶(𝑡) (4)

Where: X(t) is the source input signal, Y(t) is the receiver output signal, FFT is the fast
Fourier transform, and IFFT is the inverse fast Fourier transform.

Source Receiver Source Receiver

P-P S-S

Source Receiver

P-P Peak to peak method


S-S Start to start method
Inv Inv Inversion method

Figure 4. A representation of time-domain travel time techniques.

Once the travel times are measured, the Vs and Vp, and subsequently, the Gmax and Emax were
calculated as per Equations (5) and (6).

𝑉𝑠 = 𝐿𝑡𝑡 /𝑇𝑠 , 𝐺𝑚𝑎𝑥 = 𝜌 ∗ 𝑉𝑠2 (5)

𝑉𝑝 = 𝐿𝑡𝑡 /𝑇𝑝 , 𝐸𝑚𝑎𝑥 = 𝜌 ∗ 𝑉𝑠2 ∗ (3𝑉𝑝2 − 4𝑉𝑠2 )/(𝑉𝑝2 − 𝑉𝑠2 ) (6)

The wave path to wavelength ratio (Ltt\λ) was also calculated, and it was used as a significant
quantity to judge the accuracy of the output signal in all cases. When the ratio was less than 2, it
was not easy to calculate travel times, similar to that observed by Janusz et al. (2021) and Wang
et al. (2017).
The curves for Gmax/F(e) versus σ' are presented in Figure 5 for various travel time
measurement techniques at 95 % MDD and 80 % MDD. For dense specimens, the solid particles
are in good contact, which facilitates obtaining a clean output signal with minimal interferences
and noise. This has resulted in a smooth and almost consistent travel time measurement (Figure
5a). However, loose specimens are difficult to prepare, and the signal interpretation and analysis
are much more complex due to the interference of the p-wave part in the s-waves and the
corresponding near-field effect. This is reflected in Figure 5 (b) for an 80 % MDD dry specimen
where no two travel time techniques are identical, specifically at higher confining pressure.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 500

100 100
(a) (b)
80 80
Gmax/F(e), (MPa)

Gmax/F(e), (MPa)
60 60
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

40 40

20 s-s p-p 20 s-s inv


inv cc p-p cc1 cc2
0 0
0 50 100 150 200 250 0 50 100 150 200 250
Effective confining pressure (kPa) Effective confining pressure (kPa)

Figure 5. A typical representation of Gmax for different travel time measurement techniques
for: (a) 95 % MDD sample and (b) 80 % MDD sample.

Effect of sample preparation method and saturation

The sample preparation method influenced the shear modulus, with the values obtained with
the dry tamping technique being on the lesser side compared with the moist tamping technique
after both specimens are saturated, as can be observed from Figure 6. One reason can be the
enhancement of initial stiffness by the surface tension of water. Additionally, the shear modulus
of the saturated specimen was lower than that of the dry specimen for the same confining
pressure and sample preparation method. Both observations are consistent with those reported by
Gu et al. (2015) for natural sands. In addition, once the saturation value was reached, the p-wave
velocity remained constant at around 1650 m/s (Vp in water is ≈ 1500 m/s), which hindered the
measurement of Emax.

140
Dry tamping (DT)
120 Moist tamping (MT)
Gmax/F(e), (MPa)

100
80
60
40
20
0
0 50 100 150 200 250
Effective confining pressure (kPa)

Figure 6. Effect of sample preparation method on Gmax of coal ash

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 501

Effect of confinement and void ratio

The confinement effect on the Gmax and Emax on a sample prepared using dry tamping and
tested in its dry state at 10 kHz is shown in Figures 7 (a) and (b). The higher 'n' values indicate
the slow variation of modulus with confining pressure. However, they vary over a long range of
confining pressures. Hence, it can be observed that even though Emax varies more rapidly when
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

compared to Gmax in the lower range of confining pressure, it will not be affected by very high
confining pressures. The typical value of the fitting parameters 'A' and 'n' for natural soils was
observed to be between 7 and 14.1 and 0.4 and 0.5, respectively (Fernando and Tauta 2012).
Figure 7 (c) shows the variation of Gmax and Emax with void ratio (e). Due to the amorphous
nature of coal ash, the reported 'A' value was found to be slightly lower than expected.

250
Range of A, n values for natural soils: (a)
Upper limit, A = 14, n = 0.5
200 Lower limit, A = 7, n = 0.4
Present study (coal ash), A =7.7, n= 0.43
Gmax\F(e), (MPa)

150

100 0.43
Gmax = 7.7(σ`)

50

0
0 50 100 150 200
Effective confining pressure (kPa)
200 100
(b) E vs σ'
(c) G vs e E vs e
80
Emax \ F(e), (MPa)

150 -3.04
Gmax , Emax (MPa)

Emax = 23.5(σ')
0.37 Emax = 26.8e
60
100
40

50
20
-3.32
Gmax = 11.0e
0 0
0 20 40 60 80 100 0.5 0.7 0.9 1.1 1.3 1.5
Effective confining pressure (kPa) Void ratio (e)

Figure 7. (a) Effect of confinement on shear modulus, (b) Effect of confinement on elastic
modulus, (c) Elastic parameters as a function of void ratio

Evaluation of Poisson's ratio and bulk modulus

The Poisson's ratio (ν) was calculated using Equation (7), and the obtained value conformed
to the range of 0.15 to 0.31 in the present study.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 502

𝜈 = 0.5 ∗ (𝑉𝑝2 − 2𝑉𝑠2 )/(𝑉𝑝2 − 𝑉𝑠2 ) (7)

A general observation is that with a reduction in void ratio and an increase in density, for the
same frequency excitation and confining pressure, the Poisson's ratio of the specimen increased
toward the 0.25 - 0.30 range indicating the increase in stiffness. Sample results are shown in
Table 3. The derived elastic parameters bulk modulus (K) and Poisson’s ratio seem to be within
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

typical limits. However, there is a considerable variation in the results of elastic parameters with
the varying void ratio in the case of coal ash; hence, it becomes pertinent to consider these
variations for practical design considerations.

Table 3. Representation of evaluation of small strain elastic parameters

Void RC Travel Time Velocity (m/s) Gmax Emax ν K


ratio, (%) (ms) (MPa) (MPa) (MPa)
e Ts Tp Vs Vp
0.67 96 0.56 0.31 171.67 293.94 37.35 80.16 0.31 58.00
0.99 83 0.77 0.48 119.48 191.67 15.08 38.80 0.18 28.75
1.28 70 1.51 0.92 64.24 105.44 3.81 9.19 0.20 7.00

CONCLUSIONS

This work aimed to determine the small strain shear characteristics of coal ash and the
following are major observations:
• Based on the results of the testing and subsequent analysis, the optimal wave path to
wavelength ratio for obtaining a noise-free signal with low near-field influence is in the
range of 2 to 3. However, when the wave path to wavelength ratio was greater than 7, the
signal distortion due to damping was noticed.
• The Gmax obtained using the Start-to-Start method had less variability than other
methods.
• Signal interpretation was simpler and easier on dense specimens due to the absence of
compressional wave elements in the signal. Determining the travel time and Emax for the
p-wave was simple.
• The travel time for p-wave excitation of saturated ash is nearly constant at all applied
confining pressures, indicating the limitation of the measurement of Emax. Furthermore,
the dominant frequency range shifted above 50 kHz, resulting in p-wave excitation only
at a 50 kHz input frequency in the analysis.
• After saturation, the sample prepared with the dry tamping method had lower shear
stiffness than the one prepared with moist tamping method which is consistent with that
reported for natural sands.
• Some fitting parameters deviated slightly from the literature values for natural soils. One
reason can be attributed to the amorphous nature of coal ash.
• The bulk modulus and Poisson's ratio calculated indirectly appear to be within typical
limits. As a result, bender element testing may be an effective alternative for quick
evaluation of small strain stiffness parameters for coal ash material.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 503

REFERENCES

ASTM. ASTM D2487-17. (2019). Standard Practice for Classification of Soils for Engineering
Purposes (Unified Soil Classification System). ASTM Int., West Conshohocken, PA.
ASTM. ASTM D4318-17. (2017). Standard Test Methods for Liquid Limit, Plastic Limit, and
Plasticity Index of Soils. ASTM Int., West Conshohocken, PA.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

ASTM. ASTM D8295-19. (2019). Standard Test Method for Determination of Shear Wave
Velocity and Initial Shear Modulus in Soil Specimens using Bender Elements. ASTM
Int.,West Conshohocken.
Bachus, R. C., et al. (2019). "Characterization and Engineering Properties of Dry and Ponded
Class-F Fly Ash." J. Geotech. Geoenvironmental Eng., 145 (3): 04019003.
Choo, H., Yeboah, N. N., and Burns, S. E. (2016). "Small to intermediate strain properties of fly
ashes with various carbon and biomass contents." Can. Geotech. J., 53 (1): 35–48.
Fernando, J., and Tauta, C. (2012). "A procedure to calibrate and perform the Bender Element
Test Element." Dyna, 176 (2012): 10–18.
Gu, X., Yang, J., and Huang, M. (2013). "Laboratory measurements of small strain properties
of dry sands by bender element." Soils Found., 53 (5): 735–745. Japanese Geotechnical
Society.
Gu, X., Yang, J., Huang, M., and Gao, G. (2015). "Bender element tests in dry and saturated
sand: Signal interpretation and result comparison." Soils Found., 55 (5): 951–962.
Elsevier.
Jakka, R. S., Datta, M., and Ramana, G. V. (2010). "Liquefaction behaviour of loose and
compacted coal ash." Soil Dyn. Earthq. Eng., 30 (7): 580–590. Elsevier.
Janusz, P. A., Sørensen, K. K., Clausen, O. R., and Andresen, K. J. (2021). "Influence of sample
conditions on shear wave velocity measurements in a sedimentary stiff clay." Mar.
Georesources Geotechnol., 39 (4): 448–458. Taylor & Francis.
Lings, M. L., and Greening, P. D. (2001). "A novel bender/extender element for soil testing."
Geotechnique, 51 (8): 713–717. ICE.
Moldovan, I. D., Correia, A. G., and Pereira, C. (2016). "Bender-based G0 measurements: A
coupled numerical-experimental approach." Comput. Geotech., 73: 24–36.
Murillo, C., Sharifipour, M., Caicedo, B., Thorel, L., and Dano, C. (2011). "Elastic parameters of
intermediate soils based on bender-extender elements pulse tests." Soils Found., 51 (4): 637–
649. Elsevier.
Ogino, T., Kawaguchi, T., Yamashita, S., and Kawajiri, S. (2015). "Measurement deviations for
shear wave velocity of bender element test using time domain, cross-correlation, and
frequency domain approaches." Soils Found., 55 (2): 329–342. Elsevier.
Parhi, P. S., Balunaini, U., Sravanam, S. M., and Mauriya, V. K. (2020). "Site Characterization
of Existing and Abandoned Coal Ash Ponds Using Shear-Wave Velocity from Multichannel
Analysis of Surface Waves." J. Geotech. geoenvironmental Eng., 146 (11): 1–13.
Ramaiah, B. J., Jakka, R. S., and Ramana, G. V. (2010). "Shear Wave Velocity Measurements at
Slurry Deposited Coal Ash Ponds in Delhi, India." Proc. 6th Int. Conf. Environ. Geotech.,
(November): 578–581.
Sridharan, A., Pandian, N. S., and Prasad, P. S. (2000). "Liquid Limit Determination of Class F
Coal Ash." J. Test. Eval., 28 (6): 455–461.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 504

Wang, Y., Benahmed, N., Cui, Y. J., and Tang, A. M. (2017). "A novel method for determining
the small-strain shear modulus of soil using the bender elements technique." Can. Geotech.
J., 54 (2): 280–289.
Zand, B., Tu, W., Amaya, P. J., Wolfe, W. E., and Butalia, T. S. (2009). "An experimental
investigation on liquefaction potential and post-liquefaction shear strength of impounded fly
ash." Fuel, 88 (7): 1160–1166. Elsevier.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 505

Effect of Freezing-Thawing on Preconsolidation Pressure

Seyed Morteza Zeinali, S.M.ASCE1; and Sherif L. Abdelaziz, Ph.D., A.M.ASCE2


1
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
VA. Email: smzeinali@vt.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
VA. Email: saziz@vt.edu

ABSTRACT

This study aims to assess the impact of freezing-thawing cycles on the preconsolidation
pressure of saturated clays. The effect of elevated temperatures and the impact of freezing on the
preconsolidation pressure of clays have been investigated. However, there still exists a lack of
understanding about how one extreme temperature, such as freezing, impacts the
preconsolidation pressure on the other extreme, such as an elevated temperature. In this paper, a
modified temperature-controlled odometer is utilized to determine temperature effects on the
preconsolidation pressure. One-dimensional consolidation is first performed on two kaolinite
clay specimens, one at room temperature (20°C) and the other at an elevated temperature of
40°C. Moreover, another specimen is exposed to a temperature cycle of 20°C to –15° and to
40°C. Then, the specimen is again incrementally consolidated at 40°C. The preconsolidation
pressure for each specimen is estimated using the strain energy. We then assess the impact of the
freezing temperature applied to the last specimen on the preconsolidation pressure by comparing
the pressure between the three considered specimens. The results suggest that the sample
experiencing a freezing temperature prior to a heating stage shows a higher preconsolidation
pressure.

INTRODUCTION

The fight against the adverse impacts of climate change has motivated various novel
engineering applications. The temperature rise, as a consequence of climate change (Jungqvist et
al. 2014, NOAA 2020, Wang et al. 2020), causes thawing of permafrost and thus the instability
of infrastructures. These damages can be mitigated using thermal soil stabilization methods,
which maintain the permafrost (Doré et al. 2012, Filimonov and Vaganova 2013, Loktionov, et
al. 2022). Also, heat stabilization of the soil has been widely used to engineer the properties of
soils and for treatment of contaminated soils (Richards 1969, Alcocer and Chowdhury 1993). On
the other hand, the increasing tendency to reduce the consumption of fossil fuel is encouraging
the utilization of green sources of energy such as geothermal energy (Abdelaziz et al. 2011,
Knellwolf et al. 2011, Olgun et al. 2012, Coccia al. 2013, Jaradat and Abdelaziz 2018). Various
geotechnical applications, from thermal stabilization of soils to installing the heat exchanger
piles, subject the hosting ground to cycles of temperature changes which potentially alter the
preconsolidation pressure of the ground. Numerous studies have investigated the impact of
temperature change on mechanical properties of the soils such as shear strength (Broms and Yao
1964, Cekerevac et al. 2005, Bergado et al. 2009, Steiner et al. 2018, Jaradat and Abdelaziz
2020) and volume change (Delage and Lefebvre 1984, Abuel-Naga et al. 2005, Bai et al. 2014,
Samarakoon and McCartney 2020, Zeinali and Abdelaziz 2021). However, the effect of one

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 506

temperature extreme such as freezing, on the mechanical properties of the soil at the other
extreme, such as heating, is still unclear. This study aims to assess the impact of heating and
freezing before heating on preconsolidation of kaolinite clay.
Currently, it has been observed that the effect of increasing temperature on preconsolidation
pressure is highly dependent to stress history. While the preconsolidation pressure of highly
overconsolidated clays decreases with heating, normally consolidated clays experience an
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

increase in preconsolidation pressure with temperature (Campanella and Mitchell 1968, Plum
and Esrig 1969, Abuel-Naga et al. 2005, Kong et al. 2020). Also, heating cooling cycles increase
the pre-consolidation pressure for normally consolidated clays (Samarakoon and McCartney
2020). The location of the point, which corresponds to void ratio and stress of a sample, at the
end of heating-cooling cycles relative to the recompression or virgin compression curves
determines the evolution of preconsolidation pressure with temperature. If the void ratio of a
sample increases after heating-cooling, the preconsolidation would increase; on the contrary,
permanent contraction at the end of the thermal cycle increases to preconsolidation pressure.
Moreover, freezing induces a pseudo-overconsolidation state and increases the apparent
preconsolidation pressure (Chamberlain 1981, Qi et al. 2010). However, while some studies
suggest that the freezing-induced disturbance makes the post-thaw preconsolidation pressure
undetectable (Graham and Au 1985, Qi et al. 2006), other experimental investigation showed an
increase in preconsolidation pressure after thawing (Qi et al. 2008).
Despite the discussed efforts, the impact of the phase change (frozen to thawed and vice
versa) and thermal cycles (e.g., freezing-heating) on the consolidation characteristics of clays
needs more investigation. For this purpose, in this study a thermally modified oedometer was
utilized to assess the impact of thermal history on preconsolidation pressure of kaolinite clay.
Three samples were considered in this paper. One sample was subjected to incremental
consolidation at room temperature, as the reference sample. Afterwards, consolidation tests were
performed on two other samples at 40°C; however, among these two, one sample experienced a
freezing temperature of -15°C before being consolidated at 40°C. In the next step, the
preconsolidation pressures were estimated using work method (Becker et al. 1987).

SAMPLE PREPARATION AND EXPERIMENTAL METHODS

The clay used in this study is the Edgar Plastic Kaolin, from Edgar, Florida,in powder form.
According to United Soil Classification System (USCS), EPK is classified as highly plastic clay
(CH) with liquid and plastic limits of 67 and 32%, respectively. The results of benchtop X-ray
diffraction showed that more than 96% of minerals in EPK is Kaolnite (Darbari et al. 2017,
Jaradat et al. 2017). The bulk clay samples were prepared by mixing the clay powder with de-
ionized water at one and half times the liquid limit. The resulting slurry was then poured into a
mold and kept under the weight of the loading cap for 24 hours to avoid piping (Mitchell and
Soga 2005). After 24 hours, the slurry was subjected to one-dimensional incremental loading up
to 100 kPa. Upon the completion of the consolidation, the clay block was extracted, and the
oedometer samples were trimmed out of the block.
Table 1 lists the initial void ratio and the thermal load subjected to each sample considered in
this study prior to initiation of the consolidation test. Figure 1 presents the thermally modified
oedometer used to perform consolidation experiments at the desired temperatures. A fixed
consolidation ring is placed in the chamber. The inner coil inside the well-insulated oedometer
chamber is connected to a temperature controller unit manufactured by Julabo. This unit is able

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 507

to change and maintain the temperature inside the oedometer chamber by circulating a
temperature-controlled oil inside the internal coil. The temperature of the cell fluid is monitored
using a submerged temperature sensor, and the temperature was maintained by automatic
adjustment of the temperature of the circulating oil.

Table 1. Samples and the corresponding temperature changes.


Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Void ratio after thermal


Sample Initial Void Ratio, load, Temperature Change
ID e eth
S1 1.27 -- --
S2 1.26 1.33 20°C - 40°C
S3 1.30 1.15 20°C - (-15°C) - 40°C

Temperature
Sensor
Temperature
Contoller

Consolidation
Chamber

Figure 1. Thermally modified oedometer.

Consolidation experiment for the first sample was performed at a temperature of 20°C. For
the samples experiencing temperature changes, first, the sample is consolidated under a seating
vertical stress of 5 kPa while maintaining 20°C. Providing that the clay block is consolidated
under 100 kPa, the vertical stress of 5 kPa yields to an overconsolidation ratio (OCR) of 20.
Furthermore, once the consolidation at the initial vertical stress is completed, corresponding
temperature change is applied to the sample. Finally, after each consolidation experiment, the
preconsolidation pressures were estimated using the work method (Becker et al. 1987). This
method relies on calculating work per unit volume and plotting the cumulative work against the
effective stress. In this two-dimensional space, the data points can be approximated by two lines.
The intersection of these two lines indicates the preconsolidation pressure.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 508

RESULTS AND DISCUSSION

Figure 2 presents the thermally induced volumetric strain for samples S2 and S3. According
to Figure 2 (a), S2 sample demonstrated expansive behavior upon heating as expected (Baldi et
al. 1988, Cekerevac and Laloui 2004, Abuel-Naga et al. 2007). This means that for S2, the void
ratio at the end of the temperature change, before the initiation of incremental loading, is higher
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

than the initial value. On the other hand, the third sample expanded during freezing as a result of
the expansion of pore water, as shown in Figure 2 (b). It should be noted that the initial
contraction in Figure 2 (b) corresponds to the cooling stage during freezing, where the negative
pore pressure increases the effective stress. This observation is in accordance with previous
studies on frozen soils (Morgenstern and Nixon 1971, Chamberlain 1981). However, changing
the temperature from -15°C to 40°C induced contraction at the end of the thermal loading, e.g.,
at the end of a freezing-thawing cycle. Therefore, there remains a permanent reduction in the
void ratio.

-1.6 -6
(a) (b)
-1.4
-4
-1.2
Volumetric Strain (%)

-2
-1

-0.8 0

-0.6
2
-0.4

20°C - 40°C, S2 4
-0.2
20°C - (-15°C) - 40°C, S3

0 6
0 200 400 600 800 1000 0 500 1000 1500 2000
Time (min) Time (min)

Figure 2. Thermally-induced volumetric strains for (a) S2 and (b) S3.

Figure 3 presents the consolidation results and experimental data interpretation using the
work method. The procedure for estimating pre-consolidation pressure is also demonstrated as a
blowup on the work-stress curve in the initial application in Figure 3 (b). For the sample at room
temperature. S1, a preconsolidation pressure of 100 kPa was obtained. Moreover, the
consolidation curves for samples S2 and S3 are also presented in Figure 4. From Figure 4 (a), the
right-wise shift for the sample subjected to freezing first is observed, indicating an increase in the
preconsolidation pressure. This observation was also confirmed using the work method, as
illustrated in Figure 4(b). The preconsolidation pressure was calculated to be 78 and 118 kPa for
S2 and S3, respectively. Figure 4 (c) and 4 (d) demonstrate how the preconsolidation was
estimated.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 509

0 25
(a) (b)

5
20

10
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

15
Strain (%)

W (kJ/m3)
3

15
2.5

10 2

20
1.5

1
5
25
Room Temperature, S1 0.5

P c = 100 kPa
0
0 100 200 300 400 500
30 0
1 10 100 1000 10000 0 1000 2000 3000 4000 5000
Vertical Stress (kPa) Vertical Stress (kPa)

Figure 3. (a) Consolidation data and (b) cumulative work per unit volume obtained from
S1.

0 25
(a) (b)
5
20

10

15
Strain (%)

W (kJ/m3)

15

20
10

25 S2
S3
5
30

35 0
1 10 100 1000 10000 0 1000 2000 3000 4000
Vertical Stress (kPa) Vertical Stress (kPa)

3 2
(c) (d)
1.8
2.5
1.6

1.4
2
1.2
W (kJ/m3)

W (kJ/m3)

1.5 1

0.8
1
0.6

0.4
0.5
0.2
P c = 78 kPa
P c = 118 kPa
0 0
0 100 200 300 400 500 0 100 200 300 400 500
Vertical Stress (kPa) Vertical Stress (kPa)

Figure 4. (a) Consolidation curves for S2 and S3 and (b, c, d) preconsolidation pressure.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 510

The results from consolidation experiments revealed that the evolution of preconsolidation
with temperature depends on the previous thermal history. This change in behavior can be
analyzed using e-log(p') curves. Figure 5 shows overconsolidation recompression and virgin
compression lines at two temperatures of T1 and T2. The virgin compression line for the higher
temperature of T2 will be below but with the same slope as that of T1 (Delage, Sultan et al. 2000,
François and Laloui 2010, Mon, Hamamoto et al. 2018, Samarakoon and McCartney 2020).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

However, the experimental results in this study showed that virgin compression curve for the
sample experiencing freezing first, (T'2) is rightward compared to T1. As discussed earlier, in this
study the temperature of the oedometer samples was changed under an OCR of 20. Therefore,
before any thermal loading, both samples were at point A. S2, expanded upon heating. Thus, the
soil was at point B prior to the continuation of mechanical loading, as shown in Figure 5 (a).
With the initiation of incremental mechanical loading, the soil followed the recompression line
intersecting the T'2 virgin compression at point C, at lower effective stress than the initial
preconsolidation pressure. On the other hand, S3 expanded during freezing and contracted during
the heating stage. Therefore, the final void ratio after the thermal load applied to S3 is placed
below the initial point of A. At this point when the consolidation starts, the soil will follow the
recompression curve, reaching the virgin compression curve at point C, which indicates the
location of ultimate preconsolidation pressure after the temperature change. It is deduced that the
relative location of the void ratio before and after any temperature change determines the
evolution of preconsolidation pressure with temperature.

e e
(a) (b)

T2 T1 T T2 T1 T

B
C T2=T T1
A A
B
C

Log p Log p
P c(T2) P c(T1) P c(T1) P c(T

Figure 5. Evolution of void ratio with (a) heating and (b) freezing-heating for S2 and S3.

CONCLUSION

In this paper, the effect of heating and freezing-thawing on preconsolidation pressure of


kaolinite clay was investigated. For this purpose, a modified oedometer was utilized, and three
samples were tested. One sample was subjected to mechanical loading at 20°C. Two other
consolidation experiments were performed at 40°C. However, among these two, one sample was
first subjected to a freezing temperature of -15°C and then raised to 40°C. The consolidation data
was then interpreted using the work method. This method was chosen to minimize the extent of
the subjectivity of conventional Casagrande's method. The thermally induced volumetric strains
prior to incremental mechanical loading showed that the expected volumetric response of highly

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 511

overconsolidated clays to elevated temperatures is different when first subjected to freezing.


While the overconsolidated clay in this study expanded during heating, exposing the sample to a
frozen stage completely alters the volumetric response to contraction upon heating. Moreover,
the results revealed that the preconsolidation pressure increased with the freezing-thawing phase.
the findings revealed that the yield stress at one temperature is highly dependent on the previous
temperature that the soil experienced. A phenomenon that can be inferred as a "thermal stress
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

history", similar to conventional stress history of clays but dependent on the previous
temperatures. While more investigation is required under various temperatures and stress
histories, thermal cycles can modify the thermal response of the soils and also can be utilized for
soil improvement.

ACKNOWLEDGMENT

This material is based upon work supported by the Broad Agency Announcement Program
and the Cold Regions Research and Engineering Laboratory (ERDC-CRREL) under Contract
No. W913E522C0001. Any opinions, findings and conclusions, or recommendations expressed
in this material are those of the author(s) and do not necessarily reflect the views of the Broad
Agency Announcement Program and ERDC-CRREL.

REFERENCES

Abdelaziz, S. L., Olgun, C. G., and Martin, J. R., II. 2011. Design and operational considerations
of geothermal energy piles. In Geo-frontiers 2011: Advances in geotechnical engineering,
450-59.
Abuel-Naga, H. M., Bergado, D. T., and Bouazza, A. 2007. “Thermally induced volume change
and excess pore water pressure of soft Bangkok clay.” Eng. Geol., 89: 144-54.
Abuel-Naga, H. M., Bergado, D. T., Soralump, S., and Rujivipat, P. 2005. “Thermal
consolidation of soft Bangkok clay.” Lowland Technology International Journal, 7: 13-22.
Alcocer, C. F., and Chowdhury, H. R. 1993. Experimental study of an environmental
remediation of Gulf coast crude-oil-contaminated soil using low-temperature thermal
treatment. In SPE Western Regional Meeting. OnePetro.
Bai, B., Guo, L., and Han, S. 2014. “Pore pressure and consolidation of saturated silty clay
induced by progressively heating/cooling.” Mech. Mater., 75: 84-94.
Baldi, G., Hueckel, T., and Pellegrini, R. 1988. “Thermal volume changes of the mineral–water
system in low-porosity clay soils.” Can. Geotech. J., 25: 807-25.
Becker, D. E., Crooks, J. H. A., Been, K., and Jefferies, M. G. 1987. “Work as a criterion for
determining in situ and yield stresses in clays.” Can. Geotech. J., 24: 549-64.
Bergado, D., Liu, M. D., Abuel-Naga, H. M., and Carter, J. P. 2009. Predicting the
thermomechanical behaviour of natural clays.
Broms, B. B., and Yao, L. Y. C. 1964. “Shear strength of a soil after freezing and thawing.”
Journal of the Soil Mechanics and Foundations Division, 90: 1-25.
Campanella, R. G., and Mitchell, J. K. 1968. “Influence of temperature variations on soil
behavior.” Journal of the Soil Mechanics and Foundations Division, 94: 709-34.
Cekerevac, C., and Laloui, L. 2004. “Experimental study of thermal effects on the mechanical
behaviour of a clay.” Int. J. Numer. Anal. Met., 28: 209-28.
Cekerevac, C., Laloui, L., and Vulliet, L. 2005. “A novel triaxial apparatus for thermo-
mechanical testing of soils.” Geotech. Test. J., 28: 161-70.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 512

Chamberlain, E. J. 1981. “Overconsolidation effects of ground freezing.” Eng. Geol., 18: 97-110.
Coccia, C. J. R., Gupta, R., Morris, J., and McCartney, J. S. 2013. “Municipal solid waste
landfills as geothermal heat sources.” Renewable and sustainable energy reviews, 19: 463-
74.
Darbari, Z., Jaradat, K. A., and Abdelaziz, S. L. 2017. “Heating–freezing effects on the pore size
distribution of a kaolinite clay.” Environmental earth sciences, 76: 713.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Delage, P., and Lefebvre, G. 1984. “Study of the structure of a sensitive Champlain clay and of
its evolution during consolidation.” Can. Geotech. J., 21: 21-35.
Delage, P., Sultan, N., and Cui, Y. J. 2000. “On the thermal consolidation of Boom clay.” Can.
Geotech. J., 37: 343-54.
Doré, G., Ficheur, A., Guimond, A., and Boucher, M. 2012. Performance and cost-effectiveness
of thermal stabilization techniques used at the Tasiujaq airstrip. in, Cold Regions Engineering
2012: Sustainable Infrastructure Development in a Changing Cold Environment.
Filimonov, M. Y., and Vaganova, N. A. 2013. “Simulation of thermal stabilization of soil around
various technical systems operating in permafrost.” Appl. Math. Sci, 7: 7151-60.
François, B., and Laloui, L. 2010. “An oedometer for studying combined effects of temperature
and suction on soils.” Geotech. Test. J., 33: 112-22.
Graham, J., and Au, V. C. S. 1985. “Effects of freeze–thaw and softening on a natural clay at low
stresses.” Can. Geotech. J., 22: 69-78.
Jaradat, K. A., and Abdelaziz, S. L. 2020. “Thermomechanical Triaxial Cell for Rate-Controlled
Heating-Cooling Cycles.” Geotech. Test. J., 43.
Jaradat, K. A., Darbari, Z., Elbakhshwan, M., Abdelaziz, S. L., Gill, S. K., Dooryhee, E., and
Ecker, L. E. 2017. “Heating-freezing effects on the orientation of kaolin clay particles.” Appl.
Clay. Sci., 150: 163-74.
Jaradat, K., and Abdelaziz, S. 2018. Temperature-dependent load-displacement curves of heat
exchanger piles in sand. IFCEE 2018.
Jungqvist, G., Oni, S. K., Teutschbein, C., and Futter, M. N. 2014. “Effect of climate change on
soil temperature in Swedish boreal forests.” PloS one, 9: e93957.
Knellwolf, C., Peron, H., and Laloui, L.. 2011. “Geotechnical analysis of heat exchanger piles.”
J. Geotech. Geonviron., 137: 890-902.
Kong, L., Yao, Y., and Qi, J. 2020. “Modeling the combined effect of time and temperature on
normally consolidated and overconsolidated clays.” Acta Geotech., 15: 2451-71.
Loktionov, E. Y., Sharaborova, E. S., and Shepitko, T. V. 2022. “A sustainable concept for
permafrost thermal stabilization.” Sustainable Energy Technologies and Assessments, 52:
102003.
Mitchell, J. K., and Soga, K. 2005. Fundamentals of soil behavior (John Wiley & Sons New
York).
Mon, E. E., Hamamoto, S., Kawamoto, K., Komatsu, T., and Møldrup, P. 2018. “Temperature
effects on geotechnical properties of kaolin clay: simultaneous measurements of
consolidation characteristics, shear stiffness, and permeability using a modified oedometer.”
J. Geol. Sci., 1.
Morgenstern, N. R., and Nixon, J. F. 1971. “One-dimensional consolidation of thawing soils.”
Can. Geotech. J., 8: 558-65.
NOAA. 2020. “State of the Climate: Global Climate Report.” NOAA National Centers for
Environmental Information, State of the Climate: Global Climate Report for 2020.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 513

Olgun, C. G., Abdelaziz, S. L., and Martin, J. R. 2012. Long-Term Performance and Sustainable
Operation of Energy Piles. In Proc. International Conference on Sustainable Design,
Engineering, and Construction, 534-42.
Plum, R. L., and Esrig, M. I. 1969. “Some temperature effects on soil compressibility and pore
water pressure.” Highway Research Board Special Report.
Qi, J., Hu, W., and Ma, W. 2010. “Experimental study of a pseudo-preconsolidation pressure in
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

frozen soils.” Cold Reg. Schi. Technol., 60: 230-33.


Qi, J., Ma, W., and Song, C. 2008. “Influence of freeze–thaw on engineering properties of a silty
soil.” Cold Reg. Schi. Technol., 53: 397-404.
Qi, J., Vermeer, P. A., and Cheng, G. 2006. “A review of the influence of freeze‐thaw cycles on
soil geotechnical properties.” Permafrost Periglac., 17: 245-52.
Richards, F. 1969. “Temperature effects on the engineering properties and behavior of soils.”
Special Report: 9.
Samarakoon, R. A., and McCartney, J. S. 2020. Effect of Drained Heating and Cooling on the
Preconsolidation Stress of Saturated Normally Consolidated Clays. In Geo-Congress 2020:
Foundations, Soil Improvement, and Erosion, 620-29. American Society of Civil Engineers
Reston, VA.
Steiner, A., Vardon, P. J, and Broere, W. 2018. “The influence of freeze–thaw cycles on the
shear strength of illite clay.” Proceedings of the Institution of Civil Engineers-Geotechnical
Engineering, 171: 16-27.
Wang, Q., Qi, J., Wu, H., Zeng, Y., Shui, W., Zeng, J., and Zhang, X. 2020. “Freeze-Thaw cycle
representation alters response of watershed hydrology to future climate change.” Catena,
195: 104767.
Zeinali, S. M., and Abdelaziz, S. L. 2021. “Thermal Consolidation Theory.” J. Geotech.
Geonviron., 147: 04020147.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 514

Simple Modifications to a Direct Shear Device to Perform Constant Normal Stiffness


(CNS) Tests

C. D. P. Baxter, Ph.D., P.E., M.ASCE1; J. Fernandez Scarioni2;


A. S. Bradshaw, Ph.D., P.E., M.ASCE3; and A. Babaee4
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Dept. of Ocean, Civil, and Environmental Engineering, Univ. of Rhode Island, Narragansett, RI.
Email: cbaxter@uri.edu
2
Dept. of Ocean, Civil, and Environmental Engineering, Univ. of Rhode Island, Narragansett, RI.
Email: jfernandezs@my.uri.edu
3
Dept. of Ocean, Civil, and Environmental Engineering, Univ. of Rhode Island, Narragansett, RI.
Email: abrads@uri.edu
4
Dept. of Ocean, Civil, and Environmental Engineering, Univ. of Rhode Island, Narragansett, RI.
Email: amir_babaee@uri.edu

ABSTRACT

This paper describes a simple approach to modify a direct shear test device to perform
interface shear tests under constant normal stiffness (CNS) conditions. There is continued
interest in the use of piles and pile anchors for both fixed and floating offshore wind structures,
particularly in the sandy soils along the East Coast of the United States. Design analyses and
modeling may require CNS testing, which closely represents the boundary conditions near the
interface of the pile during axial monotonic and cyclic loading. The simple and economical
modifications allow for the testing of any pile material-sand interface by mechanically
maintaining constant normal stiffness conditions on the sample during shear.

INTRODUCTION

The offshore wind industry is moving into new regions of the world and further offshore. The
first offshore wind installation in the US at Block Island, Rhode Island utilized piled jacket
structures. Future fixed and floating platforms here and in other regions may also utilize piles
and pile anchors. In both the fixed and floating applications the piles are subjected to cyclic axial
loads may cause a reduction in the side shear stress and capacity of the pile, a term commonly
referred to as cyclic degradation (e.g., Poulos 1988). Cyclic degradation is a complex process
and depends largely on the specific pile material-soil combination as well as the cyclic load
levels and number of cycles.
There are a variety of approaches that have been proposed to assess cyclic degradation in
design (e.g. Jardine et al. 2012). One of them involves the analysis of cyclic degradation using
any one of a number of computer models that have been published in the literature (e.g. Zhou et
al. 2019) and that have been developed in-house by specialized offshore geotechnical
engineering companies. Given that site-specific pile load tests are generally not available for
design, it can be beneficial to have site-specific laboratory cyclic data to assist in the calibration
and/or validation of these models.
In sands, cyclic degradation occurs within a thin band of soil at the pile interface as the pile is
axially loaded. Boulan and Foray (1986) were the first explain the mechanics using the simple
spring model as illustrated in Figure 1.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 515
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Conceptual illustration of constrained dilation within the interface shear band
during axial cyclic loading of a pile (Lehane and White 2004).

The shear band has a thickness typically between 1 and 10 times the median grain diameter
(DeJong and Westgate 2009) and is constrained laterally by the soil outside the shear band under
constant normal stiffness (CNS) as represented by the spring. The change in normal stress is
defined by the following equation (Boulan and Forray 1986):

4G Dt
Ds = - = -K n Dt
D

where =change in normal stress, G=shear modulus of soil outside the shear band, t=change
in shear band thickness (positive for contraction), D=pile diameter, and Kn=normal stiffness. As
described in DeJong et al. (2006) cyclic axial loading of the pile tends to incrementally densify
the shear band, which unloads the spring and reduces the normal stress and friction on the sides
of the pile. Monotonic loading can contract or expand the shear band depending on the initial
density of the shear band and interface roughness.
Laboratory CNS tests have been used to study interface shear behavior for piles for many
decades (e.g. Airy et al. 1992). A variety of test configurations have been proposed but they all
have a few common features. Typically, a soil specimen is placed within a rigid interface shear
“box” that holds a sample of soil so that it can bear directly on the interface material. An initial
normal load is then applied to the interface and then the normal stiffness is maintained through
the shear phase. Some systems have used a mechanical spring to maintain CNS conditions (e.g.
Tabucanon et al. 1985) while others have utilized active control (Porcino et al. 2003; DeJong et
al. 2003; 2006; DeJong and Westgate 2009). The next section describes how direct shear
equipment can be modified at minimal cost and effort to achieve the necessary CNS boundary
conditions.

EQUIPMENT MODIFICATIONS

In this study modifications were made to a Load Trac II Shear Test device manufactured by
Geocomp Corporation of Acton, MA (Fernandez Scarioni 2019). The basic equipment consists

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 516

of a rigid split shear box, where normal and shear loads are applied using two independent servo
motors with PID controllers. The sample is sheared along a horizontal plane by moving the
bottom half of the shear box while keeping the top half stationary. Although the equipment has a
load control system it does not currently have the capability of maintaining constant normal
stiffness conditions.
Four key modifications were made to the direct shear equipment to perform the test under
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

CNS conditions, as identified in the photograph in Figure 2. The first modification was to replace
the bottom half of the shear box with an aluminum block that allowed the interface material to be
attached. This allowed the top half of the shear box containing the soil to bear on the interface
surface. However, the top half was supported by four adjustable Teflon “feet” to minimize
friction between the shear box and the bottom plate. The second modification involved the
fabrication of a bracket that prevents the top half of the shear box from moving horizontally
during the shear phase. The third modification was the installation of two vertical rods to prevent
rocking of the top shear box during the shear phase.
The fourth modification involved replacing the rigid cross beam that comes with the system,
with a more flexible aluminum beam that served as a mechanical spring (Figure 2). The beam
attaches to two vertical loading rods, which allowed a normal load to be applied through the
“spring” that was measured by the load cell.

Figure 2. Photograph showing the four key modifications made to the Shear Trac II
equipment to perform CNS testing.

The normal stiffness (Kn) in the interface problem is defined as a normal stress per unit of
normal displacement and commonly presented in units of kPa/mm. Therefore, the cross beam

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 517

was sized to achieve the desired normal stiffness. In this case the connection points made the
structural analysis more complicated because the beam restraining knobs resulted in a connection
that was neither a “pin” nor “fixed.” To insure consistency between tests the knobs were
tightened to a specified torque using a torque wrench. However, one possible future modification
could include the attachment of “knife-edge” supports on the top and bottom of the beam at each
loading rod location that would better approximate simply supported conditions.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

The following equations were derived from beam mechanics to guide the beam design:
Simply supported beam:

4E b h 3
Kn =
A s l3

Fixed-ended beam:

16E b h3
Kn =
A s l3

where E=modulus of elasticity of beam material, b=width of beam, h=height of beam,


l=unsupported length of beam, and As=cross section area of soil sample.
Two aluminum beams were fabricated having widths of 36.4 mm and heights of 6.9 and 9.7
mm, respectively. Independent calibration of the beams within the apparatus confirmed that the
normal stiffness of the beam was between fixed and simply supported conditions and a Kn of
~7Ebh3/Asl3 provided a good approximation in this case. Since there is additional compliance in
the system in addition to the beam deflection, more accurate values of Kn were also evaluated
directly from the CNS test results. This was done by plotting the normal stress against the normal
displacement over the whole test and then taking the slope of the line.

TEST PROCEDURE AND RESULTS

To assess the validity of the results, tests were performed on a Monterey sand-steel interface
under both Constant Normal Load (CNL) and CNS conditions. Samples were prepared by dry
tamping to achieve a global relative density of either ~10% or ~65%. The sample was placed in
the machine and an initial normal stress of 100, 200, 300, or 400 kPa was applied to the sample.
A small gap (less than the median grain size of the sand) was made between the plate and bottom
of the shear box using the set screws. Typically, a larger gap is used in direct shear testing to
prevent lodging of the grains into the gap at larger displacements. However, in a CNS test sand
loss would give erroneous vertical displacement results.
For the shear phase in the CNS tests the cross bar was “locked off” by setting the vertical
(normal) motor velocity control to zero. Monotonic tests gave relatively repeatable results with
peak shear and normal stresses within 20% of each other. The variability is believed to be mostly
associated with the sample preparation rather than the test equipment. The cyclic tests were
performed at constant displacement amplitudes ranging from 0.25 mm to 2 mm. However, after
analysis of the results, the 2 mm displacement tests were not included because, for the loose
samples, the normal stress decreased to zero within the first cycle of loading.
Typical monotonic results are shown in Figure 3. As shown in Figure 3a the peak shear stress
was reached at displacements of about 0.3 to 0.5 mm. Since the normal stress changes were

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 518

small up to the peak shear stress, the CNS and CNL gave comparable shear strengths. However,
at larger post-peak shear displacements the normal stress decreased in the CNS tests particularly
in the loose sand due to shear band contraction, resulting in lower shear resistance. The
remaining test results, which are not provided here, showed an increase in the amount of
contraction with an increase in initial normal stress, which is consistent with critical state soil
mechanics.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Typical monotonic CNS and CNL test results for the Monterey sand-steel
interface.

The shear stress and normal stress at failure were used to estimate the peak interface friction
angles (’) as shown in Figure 4. As shown in the figure both the CNS and CNL tests yielded
similar interface friction angles of around 17 degrees for the loose samples and 23 degrees for
the dense samples. The results are approximately 2/3 of the peak friction angles (’) measured in

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 519

direct shear tests on the same soil and at the same relative density and normal stress conditions.
The 2/3’ rule of thumb is frequently used to estimate interface friction angle of soil-steel
interfaces, giving further confidence in the results.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Summary of peak interface friction angles evaluated from the CNS and CNL test
results on the Monterey sand-steel interface.

Typical cyclic CNS test results are shown in Figure 5. The behavior is consistent with the
behavior of “smooth” interfaces (Mortara et al. 2007; Mortara et al. 2010) where, for a given half
cycle, the normal displacement rapidly contracts (Figure 5c) and then levels out with further
shear displacement. The normal displacement accumulates with each subsequent cycle gradually
unloading the spring and reducing the normal stress.

Figure 5. Typical cyclic CNS results for the Monterey sand-steel interface.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 520

The stress paths of all tests are shown in Figure 6. The normal stress incrementally moves to
the left until the shear stress intersects with the strength envelope after which it follows the
strength envelope downward toward the origin. The cyclic tests induce cyclic shear
displacements that have the effect of reorienting the particles as would occur in a large
displacement monotonic test. As a result, the friction angle evaluated from the cyclic tests should
be comparable to a critical state interface friction angle. A value of about 16 degrees was
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

interpreted from the results for both the loose and dense sands, which was expectedly lower than
the peak interface friction angle presented earlier.

Figure 6. Compilation of cyclic stress paths from all test results on the Monterey sand-steel
interface.

CONCLUSIONS

This study presented a way to modify standard direct shear test equipment to perform
interface tests under Constant Normal Stiffness (CNS) conditions. Four modifications were made
to the equipment including (1) a base plate for attaching the interface material, (2) a bracket to
keep the top shear box stationary during the shear phase, (3) vertical posts to reduce rocking of
the top shear box during the shear phase, and (4) a flexible beam to maintain CNS conditions on
the sample during the shear phase. Tests were performed on both loose and dense Monterey sand
to provide a means to assess the validity of the results obtained using the modified equipment.
The monotonic CNS results yielded similar strengths as conventional constant normal load

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 521

(CNL) tests at the small shear displacement range. The cyclic CNS results showed a behavior
that was consistent with similar results found in the literature for smooth interfaces. Further
improvements could include active control of the normal stiffness in lieu of a mechanical spring.

ACKNOWLEDGEMENTS
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

This work was partially supported by the U.S. Bureau of Ocean and Environmental
Management (BOEM). Thanks to Fred Pease for assistance with the equipment modifications.
The second Author performed the CNS testing as part of his MS degree through the International
Engineering Program (IEP) between URI and the Technical University Braunschweig.

REFERENCES

Airy, D., Al-Douri, R. H., and Poulos, H. G. (1992). “Estimation of Pile Friction Degradation
from Shearbox Tests.” ASTM Geotechnical Testing Journal, 15(4), 388-392.
Boulon, M., and Foray, P. (1986). “Physical and Numerical Simulation of Lateral Shaft Friction
Along Offshore Piles in Sand,” Proceedings, Third International Conference on Numerical
Methods in Offshore Piling, Nantes, France, published by Editions Technip, Paris, France,
127-147.
DeJong, J. T., White, D. J., and Randolph, M. F. (2006). “Microscale Observation and Modeling
of Soil-Structure Interaction Behavior Using Particle Image Velocimetry.” Soils and
Foundations, 46(1), 15-28.
DeJong, J. T., and Westgate, Z. J. (2009). “Role of Initial State, Material Properties, and
Confinement Condition on Local and Global Soil-Structure Interface Behavior.” Journal of
Geotechnical and Geoenvironmental Engineering, 135(11), 1646-1660.
Fernandez Scarioni, J. (2019). Cyclic Constant Normal Stiffness Tests on Sand. MS Thesis,
Department of Civil and Environmental Engineering, University of Rhode Island.
Jardine, R. J., Puech, A., and Andersen, K. H. (2012). “Cyclic Loading of Offshore Piles:
Potential Effects and Practical Design.” Society of Underwater Technology, SUT-OSIG-12-
06, 59-97.
Lehane and White. (2004). “Lateral Stress Changes and Shaft Friction for Model Displacement
Piles in Sand.” Canadian Geotechnical Journal, 42(4).
Mortara, G., Mangiola, A., and Ghionna, V. N. (2007). “Cyclic Shear Stress Degradation and
Post-Cyclic Behaviour from Sand-Steel Interface Direct Shear Tests.” Can. Geotech. J.,
44(7), 739–752.
Mortara, G., Ferrara, D., and Fotia, G. (2010). “Simple Model for the Cyclic Behavior of Sand-
Steel Interfaces.” Journal of Geotechnical and Geoenvironmental Engineering, 136(7), 1004-
1009.
Porcino, D., Fioravante, V., Ghionna, V. N., and Pedroni, S. (2003). “Interface Behavior of
Sands from Constant Normal Stiffness Direct Shear Tests.” Geotech. Test. J. 26 (3): 289-301.
Poulos, H. G. (1988). “Cyclic Stability Diagram for Axially Loaded Piles.” Journal of
Geotechnical Engineering, 114(8), 877-895.
Rimoy, S. P., Jardine, R. J., and Standing, J. R. (2013). “Displacement Response to Axial
Cycling of Piles Driven in Sand.” Proceedings of the ICE – Geotechnical Engineering.
166(GE2), 131-146.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 522

Tabucanon, J. T., Airey, D. W., and Poulos, H. G. (1985). “Pile Skin Friction in Sands from
Constant Normal Stiffness Tests.” ASTM Geotechnical Testing Journal, 18(3), 350-364.
Zhou, W., Wang, L., Guo, Z., Liu, J., and Rui, S. (2019). “A Novel t-z Model to Predict the Pile
Responses Under Axial Cyclic Loadings.” Computers and Geotechnics, 112, 120-134.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 523

Assessment of US Frost Depth Maps Considering Climate Change Effects

Behrooz Daneshian, S.M.ASCE1; and Sherif L. Abdelaziz, Ph.D., A.M.ASCE2


1
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
VA. Email: behroozdaneshian@vt.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
VA. Email: saziz@vt.edu

ABSTRACT

The goal of this paper is to determine the accuracy of US frost depth maps used to determine
the foundation depth in cold regions. Estimating the frost depth continues to represent a
challenging task to engineers, specifically in cold regions. Frost depth and heave put significant
risks on several infrastructure including highways, runways, pipelines, and buildings. It is,
therefore, recommended to locate the foundation depth of this critical infrastructure below the
expected frost depth at any given location. In the United States, predictions of the frost depth for
foundation design have been performed using maps that were developed decades ago prior to the
significant climate changes we experience nowadays. In this study, we will revisit these maps
with the sole aim of assessing their accuracy considering recent weather trends. We employ
existing semi-empirical and fully empirical predictive models to estimate the frost depth at
different US locations using recent weather data. These models estimate the frost depth based on
correlations between frost depth and the cumulative freezing degree day (CFDD); in this study,
the latter is estimated using recent weather data. We then assess the accuracy of existing frost
depth maps by comparing the predicted frost depth for the selected locations to those estimated
from existing maps. The results of the analysis show that frost depth is a function of
meteorological parameters and soil’s dry density and water content. We also conclude that the
existing contour maps overestimate the freezing depth when compared to the values predicted by
the modified Berggren formula as a representative of semi-empirical equations widely employed
to calculate frost depth.

INTRODUCTION

The frost depth or freezing depth is the depth at which the groundwater in soil freezes.
Various parameters affecting the frost depth are soil thermal properties, soil water content as
well as meteorological parameters such as the average ambient temperature, wind speed, and
precipitation (Rajaei and Baladi 2015, Bowen et al. 2021). In geotechnical engineering, the exact
estimation of frost depth is of great importance as it can be considered an appropriate index
warning geotechnical engineers to build up the foundations below the frost line to minimize the
effect of frost heave of infrastructures. The significance of the frost depth can be well realized
when building code in pavement engineering is examined as most state highway agencies use
non-frost-susceptible soils to diminish the effect of frost depth and frost heave. Generally, frost-
susceptible soils are those that are able to suck up groundwater above the water table due to the
capillary effect and this occurs when the void ratio of the soil is small enough to pull up water
(Lein et al. 2019). Hence, we can conclude that coarse-grain soils such as sand and gravel are
less frost-susceptible in comparison to silts and clays in which water flows above the water table
is strongly dominated by the capillary effect.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 524

Frost depth can be estimated from numerical modeling (e.g., finite element and finite
difference) or simple empirical and semi-empirical equations. In numerical methods, the
transient heat flow within the soil layers is modeled so that the frost depth can be estimated more
precisely by considering the fact that soil’s thermal properties change as the soil start to freeze.
This is mainly because the volume fractions of water and ice would consistently change as the
freezing process begin and consequently the soil’s total thermal properties, which are a function
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

of component’s thermal properties also vary. Hence, if these methods are capable of capturing
the soil’s thermal properties more robustly, frost depth can be predicted accurately. For instance,
Lein et al. (2019) developed a thermodynamic FEM capable of changing the soil’s thermal
properties as it undergoes freezing and thawing process based on the simulation of real
experiments that were done on a range of frost-susceptible soils to estimate the frost depth.. In a
very recent study performed by Roustaei et al. (2022), a numerical model is developed and
calibrated with temperature measurements from an instrumented section of railway track in
Ontario, Canada, with which the effect of air temperature and the amount of snowfall on frost
penetration within the railway foundation are investigated. The results show that the environment
temperature has a greater effect on frost penetration than the snow cover, which is consistent
with the findings of previous studies (Erlingsson and Saliko 2020, Zhao et al. 2020, and 2021).
However, such numerical methods are handy only for a site-specific analysis, i.e., they cannot be
used for first-order estimates of the frost depth.
In addition to numerical models, there are several empirical and semi-empirical equations
used to estimate the frost depth among which Neumann equation, Stefan equation, and Modified
Berggren equation are the most-used ones. Neumann purposed Equation 1 to estimate the frost
depth by assuming one-dimensional heat transfer in a semi-infinite region that having initial
temperature above the freezing temperature. During the freezing process, the surface temperature
falls below the freezing temperature and as a result freezing propagates within the soil stratum
(Jiji 2009).

𝑝 = 𝜇√4𝑎𝑓 𝑡 (1)

where p = frost depth (m), μ = constant, ɑ = thermal diffusivity (m2/s), f = frozen, and t = time
since the start of freezing. The ɑ and μ parameters can be calculated using Equations 2 and 3.

𝜇 2 𝑎𝑓
⁡(−𝜇 2 ) 𝑎𝑓 𝑘𝑢 𝑇𝑔 exp⁡(− 𝑎𝑢 ) √𝜋𝜇𝑙
exp −√ = (2)
𝑒𝑟𝑓𝜇 𝑎𝑢 𝑘𝑓 𝑇𝑠 𝑎𝑓 𝜇 𝑐𝑝,𝑓 (𝑇𝑠 )
1 − erf⁡(√ 𝑎 ⁡)
𝑢

𝑘𝑓 𝑘𝑢
𝑎𝑓 = ⁡𝑎𝑛𝑑⁡𝑎𝑢 = (3)
𝜌𝑐𝑝 𝜌𝑐𝑝

where erf = Gauss error function, k = soil’s thermal conductivity [W/(⁰C.m)], T g = initial ground
temperature [⁰C], u = unfrozen, Ts = surface temperature [⁰C], L = latent heat of fusion [J/kg], 
= density [kg/m3], and cp = specific heat at constant pressure [J/ (kg. ⁰C)].

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 525

In 1891, the Neumann equation was modified by Stefan with the assumption of no heat
transfer in liquid phase within the soil layers. Equation 4 shows the Stefan equation to calculate
the frost depth (Jiji 2009).

2𝑘𝑓 𝑇𝑠 𝑡
𝑝=√ (4)
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

𝜌𝑙

All parameters in Equation 4 are defined similar to those for Neumann equation. Stefan
further modified his early equation by introducing two new terms to the previous equation. The
cumulative freezing index (CFI) was defined as the multiplication of time (t) to surface
temperature (Ts). Moreover, the air temperature was converted to surface temperature by using a
dimensionless multiplication parameter (n). Substituting the new defined terms to the previous
equation and adopting the English system, Stefan upgraded the equation as follows (Bianchini
and Gonzalez 2012):

48𝑘𝑓 ∗ 𝑛 ∗ 𝐶𝐹𝐼 (5)


𝑝=√
𝐿

𝐿 = 144𝑤𝛾𝑑 (6)

where p = frost depth (ft), kf = thermal conductivity of frozen soil, [Btu/ (ft. h. ⁰F)], n =
dimensionless parameter converting air temperature to surface temperature, L = volumetric latent
heat of fusion [Btu/ (ft3)], w = water content, and γd = dry density [lb/ft3]
As Stefan equation does not take volumetric heat capacity of soil and water into account, it
leads to inaccurate results. As a result of the deficiencies of the Stefan formula, the Modified
Berggren formula was developed (Aldrich and Paynter 1953). The current study focuses on
employing the Modified Berggren formula in order to calculate the frost depth at selected
locations in the US: Minnesota, Montana, and Wisconsin with the objective of investigating the
accuracy of the current contour maps estimating the maximum frost depth across the US. Due to
global warming, cold seasons are not as cold as previous years and as a result, the current frost
depth contour maps may need to be up to date. We are also of the opinion that if updated
metrological data is utilized in frost depth calculation, more accurate results will be achieved. It
should be noted that the main reason for choosing the Modified Berggren equation to calculate
frost depth is that it is widely used in the field of geotechnical and pavement engineering and
almost all shortcomings existing in previous empirical and semi-empirical models have been
addressed in this model. Being capable of capturing inherent soil properties such as an average
thermal conductivity of frozen and unfrozen parts of the soil as well as weather conditions,
Modified Berggren equation is thought to be an appropriate tool for an initial examination of the
existing frost depth contour maps.

MODIFIED BERGGREN FORMULA

Equation 7 illustrates the Modified Berggren formula. The model assumes the soil to be a
semi-infinite mass with uniform properties and uniform initial temperature (Ti) undergoing

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 526

freezing process when the surface temperature suddenly drops from Ti to below freezing
(Aldrich and Harl 1956). The Modified Berggren equation is almost the same as Stefan formula
except that the former does consider the effects of temperature changes in the soil mass by
considering a correction factor called “λ”. This correction factor is a function of two
dimensionless thermal parameters: thermal ratio (ɑ) and fusion parameter (μ). The thermal ratio
is defined as the ratio of the initial temperature differential (mean annual temperature - 32⁰) to
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the average temperature differential. Fusion is a function of the average volumetric heat capacity,
volumetric heat fusion, and average temperature differential. Figure 1 shows a plot issued by the
US Army Corps of Engineers to estimate the λ coefficient.

Figure 1. The correction coefficient (λ) vs. fusion parameters (μ) provided by the US Army
Corps of Engineers (Technical Manual Arctic And Subarctic Construction General
Provisions 1987)

48𝑘𝑛𝐹𝐼 (7)
𝑥 = 𝜆√
𝐿

where x = frost depth (ft), k = thermal conductivity of soil [Btu/ (ft. h. ⁰F)] (from Equation 8-10),
n = conversion factor from air index to surface index, FI = annual cumulative freezing index
[degree-days], L = volumetric latent heat of fusion [Btu/ft3] (from Equation 6), and λ = the
correction factor considering the effects of temperature changes, obtained from Figure 1.
Equation 8 and 9 estimate the thermal conductivity of frozen and unfrozen silt-clay soils
respectively, which are a function of soil’s dry density and water content (Farouki 1981). It is
noteworthy to mention that as the silt-clay soils are the most frost-susceptible, the authors

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 527

assumed soils in this study to be silt-clay so that maximum frost depth can be predicted. This can
be justified since the estimated frost depth from the Modified Berggren formula is supposed to
be compared with maximum frost depth value shown by the contour maps. The thermal
conductivity required to be employed in the Equation 7 is obtained by averaging the frozen and
unfrozen thermal conductivity of soil as shown in Equation 10.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

𝑘𝑓 = 0.0833[0.01(100.022𝛾𝑑 ) + 0.085(100.008𝛾𝑑 )𝑤] (8)

𝑘𝑢 = 0.0833(0.9𝐿𝑂𝐺𝑤 − 0.2)(100.01𝛾𝑑 ) (9)

𝑘𝑓 + 𝑘𝑢
𝑘= (10)
2

where kf = thermal conductivity of frozen soil [Btu/ (ft. h. ⁰F)], ku = thermal conductivity of
unfrozen soil [Btu/ (ft. h. ⁰F)], γd = soil’s dry density [lb/ft3], and w = soil’s water content (%).
Equation 11 and 12 are used to calculate thermal ratio () and fusion parameter (μ) respectively.

𝑀𝐴𝑇 − 32
𝑎=| |∗𝑡 (11)
𝑛𝐹𝐼

𝑛𝐹𝐼 𝐶𝑎𝑣𝑔
𝜇= ∗ (12)
𝑡 𝐿

where MAT = mean annual temperature [⁰F], t = length of freezing season [days], and C avg =
average volumetric heat capacity of the soil [Btu/ (ft3. ⁰F)] (from Equation 13-15).
Equation 13 and 14 are used to calculate the volumetric heat capacity for frozen and unfrozen
soils respectively. In order to calculate fusion parameter (from Equation 12), the average
volumetric heat capacity of the soil (Cavg) is calculated from Equation 15, which is the average of
volumetric heat capacity for frozen and unfrozen soils.

𝐶𝑓 = 𝛾𝑑 (0.17 + 0.5𝑤) (13)

𝐶𝑢 = 𝛾𝑑 (0.17 + 𝑤) (14)

𝐶𝑎𝑣𝑔 = 𝛾𝑑 (0.17 + 0.75𝑤) (15)

RESULTS AND DISCUSSION

Collecting weather data. In this study, the National Oceanic and Atmospheric
Administration (NOAA) database is used to access weather data corresponding to the past thirty
years for 3 different weather stations located at Minnesota, Montana, and Wisconsin as
representative cold regions in the US. Data includes daily minimum temperature, daily maximum
temperature and the geographical coordinate. The latter is of great importance to locate the
weather station on the frost depth contour maps. Only weather stations having continues weather

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 528

data must be counted on to acquire data as they will be fed to step 3 to calculate the cumulative
freezing index (CFI) and the mean annual temperature (MAT), which play essential roles in frost
depth calculation.
Recognizing the three coldest winters in the past thirty years. Daily average temperatures
are calculated based on the daily minimum and maximum temperatures from September of a
year to the end of February of the next year. Then, the average winter temperature during this
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

period is calculated and the years with minimum average temperatures are selected as the 3
coldest winters in the recent 30 years. For estimation of cold season length (t) corresponding to
the each chosen coldest winter, the number of days in that whole year with average daily
temperatures less than 32 °F is considered as the cold season length. The coldest winters, their
lengths (t), and the MAT for each weather station in this study are tabulated in Table. 1.

Table 1. The coldest winters and their corresponding average temperatures and lengths (t)
for each weather station in this study

Duluth International Airport, MN US


The coldest winters Winter average temperature (°F) t (days)
1993-1994 22.6 136
1995-1996 24.3 159
2013-2014 24.1 148
Armells Creek, MT US
1992-1993 31.3 92
1996-1997 30.5 108
2018-2019 32.8 92
Milwaukee Mitchell Airport, WI US
1995-1996 35.6 103
2013-2014 34.2 110
2014-2015 35.8 94

Cumulative freezing index (CFI) calculation for the three coldest years. The freezing
index can be determined by the summation of the degree-days from average daily temperatures.
In this method, the number of degree-days for the day is Tavg. – 32, where Tavg. is the average
daily temperature. The summation of degree days for a freezing season results in the freezing
index. This method is employed is this study to calculate the freezing index because it performs
daily basis calculation, which is more accurate in comparison to other monthly-based methods.
Table 2 illustrate the calculated cumulative freezing index (CFI) and mean annual temperature
(MAT) for the selected coldest years for each weather station.
Frost depth calculation using Modified Berggren formula. Frost depth is calculated and
then compared to the frost depth value shown by the contour maps (Figure 2) from NOAA
(1978) issued in Unified Facilities Criteria (UFC) 3-220-10, Soil Mechanics. These depths are
for extremely cold winters without much snow cover. Snow cover, especially early in the winter,
will decrease the frost depth significantly. It should be also noted that the values of frost depth
displayed on the contours in Figure 2 are actually the average frost depth maximum, which
obtained through the measurement in various regions. Investigating the parameters required for
frost depth calculation using Modified Berggren formula, it is realized that except meteorological
parameters, all other parameters are function of soil’s dry density and soil’s water content. Thus,
a parametric analysis is performed in order to figure out the effects of each variable on the frost
depth. A reasonable range for each variable is considered. For silt-clay soils a typical range for

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 529

the dry density and water content is 85 to 115 lb/ft3 and 15 to 35 %, respectively (Goldsmith et
al. 2001). Two case scenarios are assumed for the parametric analysis in this study. In case 1, a
15% water content is assumed and dry density was varied between 85 to 115 lb/ft3, whereas in
case 2, a 100 lb/ft3 dry density was assumed and the water content was varied between 15 to 35
%. Frost depth versus dry density and water content are illustrated in Figure 3. As can be seen,
the freezing depth decreases with increasing the water content and increases with the dry density.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Additionally, it seems that frost depth variation due to changes in the dry density is slightly
higher compared to changes in the water content.

Table 2. The coldest winters and their corresponding cumulative freezing index (CFI) and
mean annual temperature (MAT) for each weather station in this study

Duluth International Airport, MN US


The coldest winters Cumulative freezing index (CFI) MAT
1993-1994 -2701.5 37.2
1995-1996 -2971.5 36.2
2013-2014 -3235 36.8
Armells Creek, MT US
1992-1993 -1625 43.3
1996-1997 -1875 43.5
2018-2019 -1556 43.9
Milwaukee Mitchell Airport, WI US
1995-1996 -1230 45.3
2013-2014 -1631 44.8
2014-2015 -1166 46.3

Figure 2. Maximum depths (in meters) of frost penetration in the US (NOAA 1978).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 530

(a) 7.5 (b) 7.5

MN
7 7
MT
6.5 WI 6.5

Frost depth (ft) 6 6

5.5 5.5
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

5 5

4.5 4.5

4 4

3.5 3.5

3 3
85 95 105 115 15 20 25 30 35
Dry density (lb/ft3) Water content (%)

Figure 3. Frost depth calculation performed in parametric analysis a) frost depth vs. dry
density, b) frost depth vs water content.

Table 3 demonstrates statistical parameters carried out on the results of parametric analysis.
As is shown, except for Duluth International Airport, MN US for which maximum (Max.) value
of the frost depth corresponding the case 1 slightly exceeds the frost depth value shown in
contour map, for other stations, all minimum (Min.), Max., and average (Avg.) values in both
cases are less than values estimated from contours implying that the current contour map
overestimates the freezing depth. Moreover, the standard deviations (Std. deviation) calculated
for case 1 are higher than those calculated for case 2 indicating the fact that frost depth is more
sensitive to dry density variations than changes in the water content. In order to investigate the
effect of climate change on frost depth due to global warming, frost depth is calculated in 3-year
period starting from 1991 to 2021 (Figure 4). Although fluctuations are obvious in these plots,
the overall trends seem to be downward expressing the fact that global warming has increased
the average annual temperature and winters are not as cold as the past. Accordingly, calculating
the frost depth based on the cold winters in the past 30 years might be extremely conservative
and unrealistic.

Table 3. Statistical parameters performed on the results of parametric analysis for frost
depth estimation * All statistical parameters (i.e., Min., Max., Avg., Std. deviation)
tabulated in table are in feet.

Duluth International Airport, MN US


Value in
Case Min. Max. Avg. Std. deviation Avg.-Std. deviation Avg.+Std. deviation
contour
1 5.89 7.2 6.49 0.47 6.02 6.97
6.56
2 5.77 6.45 6.07 0.27 5.80 6.34
Armells Creek, MT US
1 4.06 4.96 4.47 0.33 4.15 4.80
5.33
2 4.04 4.44 4.22 0.16 4.06 4.38
Milwaukee Mitchell Airport, WI US
1 3.47 4.24 3.82 0.28 3.55 4.10
4.51
2 3.50 3.80 3.63 0.12 3.51 3.75

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 531

5.5

5 MN

MT
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

4.5

Frost depth(ft)
WI

3.5

2.5

2
91-94
94-97
97-00
00-03
03-06
06-09
09-12
12-15
15-18
18-21
Date (3-year Period)

Figure 4. Frost depths calculated in 3 year-periods for each weather station

CONCLUSIONS

The findings of the current study can be summarized as follows:


• Frost depth is a function of both metrological parameters and soil properties. Soil
properties involved in frost depth calculation include soil thermal properties (e.g., soil’s
thermal conductivity, volumetric latent heat of fusion, and fusion parameter) are
functions of soil’s dry density and water content. Increasing water content causes
freezing depth to decrease, while a rise in dry density increases it. Besides, the variation
in frost depth caused by a change in dry density seems to be higher than that induced by a
change in water content.
• As frost depths estimated by the Modified Berggren formula with updated meteorological
parameters were less than the values obtained from current contours, one may conclude
that the contour map needs to be upgraded due to their frost depth overestimations. This
can be justified by the fact that the average annual temperature is rising due to global
warming and winters are no longer as cold, resulting in a reduction in frost depth
compared to the past.

AKNOWLEDGEMENT

This material is based upon work supported by the Broad Agency Announcement Program
and the Cold Regions Research and Engineering Laboratory (ERDC-CRREL) under Contract
No. W913E522C0001. Any opinions, findings and conclusions or recommendations expressed in
this material are those of the author(s) and do not necessarily reflect the views of the Broad
Agency Announcement Program and ERDC-CRREL.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 532

REFERENCES

“Arctic and Subarctic Construction Calculation Methods for Determination of Depths of Freeze
and Thaw in Soils”. (1966). TM 5-852-6. U.S. Army, U.S. Air Force.
Aldrich, H. P., and Paynter, H. M. (1953). “Analytical Studies of Freezing and Thawing of
Soils”. Technical Report No. 42. U.S. Army Corps of Engineers, Arctic Construction and
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Frost Effects Laboratory, Boston, Mass.


Aldrich, H. P., Jr. (1956). “Frost penetration below highway and airfield pavements”. Highway
Research Board Bulletin, 135.
Bianchini, A., and Gonzalez, C. R. (2012). “Pavement-transportation computer assisted
structural engineering (PCASE) implementation of the modified Berggren (ModBerg)
equation for computing the frost penetration depth within pavement structures”. ERDC/GSL-
TR-12-15. US Army Corps of Engineers, Engineer Research and Development Center,
Vicksburg, Miss.
Erlingsson, S., and Saliko, D. (2020). “Correlating air freezing index and frost penetration
depth—a case study for Sweden”. In Proceedings of the 9th International Conference on
Maintenance and Rehabilitation of Pavements—Mairepav9, 847-857. Springer, Cham.
Farouki, O. T. (1981). Thermal properties of soils. Cold Regions Research and Engineering Lab
Hanover NH.
Goldsmith, W., Silva, M., and Fischenich, C. (2001). Determining optimal degree of soil
compaction for balancing mechanical stability and plant growth capacity. Engineer Research
and Development Center Vicksburg Ms Environmental Lab.
Jiji, L. M. (2009). Heat convection. Springer Science & Business Media, Berlin.
Lein, W. A., Slone, S. M., Smith, C. E., and Bernier, A. P. (2019). “Frost Depth Penetration and
Frost Heave in Frost Susceptible Soils”. In Airfield and Highway Pavements 2019: Testing
and Characterization of Pavement Materials, ASCE, Reston, VA, 493-503.
NOAA (National Oceanic and Atmospheric Administration). (1978). Geodetic Bench Marks.
U.S. Department of Commerce - National Oceanic and Atmospheric Administration.
Rajaei, P., and Baladi, G. Y. (2015). “Frost depth: general prediction model”. Transportation
Research Record, 2510(1), 74-80.
Roustaei, M., Hendry, M. T., and Roghani, A. (2022). “Investigating the mechanism of frost
penetration under railway embankment and projecting frost depth for future expected
climate: A case study”. Cold Regions Science and Technology, 197, 103523.
Straub, A. L., and Wegmann, F. J. (1965). “Determination of Freezing Index Values”. Highway
Research Record, 68, 17-30.
Tai, B., Yue, Z., Sun, T., Qi, S., Li, L., and Yang, Z. (2021). “Novel anti-frost subgrade bed
structures a high speed railways in deep seasonally frozen ground regions: Experimental and
numerical studies”. Construction and Building Materials, 269, 121266.
Zhao, X., Zhang, H., Lai, H., Yang, X., Wang, X., and Zhao, X. (2020). “Temperature field
characteristics and influencing factors on frost depth of a highway tunnel in a cold region”.
Cold regions science and technology, 179, 103141.
Zhou, J., Zhao, W., and Tang, Y. (2021). “Practical prediction method on frost heave of soft clay
in artificial ground freezing with field experiment”. Tunneling and Underground Space
Technology, 107, 103647.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 533

Experimental Investigation on Thermal and Electrical Properties of Binary Soil Mixtures

Gaby Vasquez1; Liang Li2; and Hoyoung Seo, Ph.D., P.E., M.ASCE3
1
Staff Engineer, Terracon, Dallas, TX. Email: gaby.vasquez@terracon.com
2
Ph.D. Candidate, Dept. of Civil, Environmental, and Construction Engineering, Texas Tech
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Univ., Lubbock, TX. Email: liang.li@ttu.edu


3
Associate Professor, Dept. of Civil, Environmental, and Construction Engineering, Texas Tech
Univ., Lubbock, TX. Email: hoyoung.seo@ttu.edu

ABSTRACT

This study investigates the effect of the percentage (f) of finer particles on thermal and
electrical properties of binary soil mixtures via well-controlled laboratory tests. Binary mixtures
of a fine silica sand (Ottawa F-55 sand) and a medium-coarse sand (ASTM 20/30 sand) were
deposited at eight different mixing ratios in a custom-built soil box via air pluviation. After
deposition, a thermal needle was inserted into the dry mixture sample to measure thermal
conductivity. The thermal needle was then removed, and the mixture sample was saturated with a
0.1% NaCl solution. After completion of the saturation process, electrical resistivity and thermal
conductivity of the saturated mixture sample were measured using a Nilsson meter and a thermal
needle, respectively. Results from a series of laboratory tests showed that the void ratios of the
binary soil mixtures initially decreased with the increasing percentage of finer particles and
achieved the densest condition at f = 30%, but further increase in f led to an increase of void
ratio. Both thermal conductivity and electrical resistivity increased as f increased, peaked at
about f = 30%–50%, and then decreased. Also, the thermal conductivity in the saturated
condition was about 8.3 times that in the dry condition.

INTRODUCTION

Thermal and electrical properties of soils are critical parameters in many industrial
applications, such as geothermal energy foundations (Laloui et al. 2006), high-voltage buried
power cables (Cao et al. 2021), oil and gas pipelines (Slegel and Davis 1977), thermal solar
energy storage facilities (Brosseau et al. 2005), nuclear waste disposals (Tang et al. 2008), and
site characterization (Sudha et al. 2009). Previous research studies have shown that thermal
conductivity and electrical resistivity of saturated soils increase as soils become denser
(Friedman 2005; Wallen et al. 2016). Also, it has been known that the thermal conductivity
increases with increasing water content whereas the electrical resistivity decreases with
increasing water content (De Vries 1963; Friedman 2005; Tong et al. 2016). There have been
further efforts to establish relationship between thermal conductivity and electrical resistivity
(Sreedeep et al. 2005; Wang et al. 2017; Sun and Lu 2019).
Despite these efforts, thermal and electrical properties of soils have received less attention in
geotechnical engineering compared to mechanical properties of soils such as compression
characteristics and shear strength. Most of the previous studies mainly focused on identifying
key variables that affect thermal and electrical behavior of homogenous soils. Natural soils are
often mixtures of two or more different-sized soils, but high-quality data on thermal and
electrical properties of mixtures of various soil types are scarce in the literature. In this study, the

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 534

authors investigated the effect of the percentage (f) of finer particles on thermal and electrical
properties of binary soil mixtures via well-controlled laboratory tests.

EXPERIMENTAL METHODS

Test Soils. Binary soil mixtures in this study were prepared by mixing a fine silica sand
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(commercial name: Ottawa F-55) and a medium-coarse sand (commercial name: ASTM 20/30)
at eight different, fine-to-coarse mixing ratios based on weight — 0:10, 1:9, 2:8, 3:7, 4:6, 5:5,
8:2, and 10:0 (i.e., the percentage of finer particles f = 0, 10, 20, 30, 40, 50 80, and 100%,
respectively). Table 1 summarizes properties of the test soils, and Figure 1 presents particle size
distribution curves of the eight binary mixtures, including f = 0 and 100%, prepared in this study.

Table 1. Properties of test soils

Soil Type
Properties
Ottawa F-55 sand ASTM 20/30 sand
Effective size, D10 (mm) 0.157 0.611
Particle size corresponding to 30%
0.211 0.669
passing, D30 (mm)
Mean particle size, D50 (mm) 0.256 0.727
Particle size corresponding to 60%
0.276 0.756
passing, D60 (mm)
Min. void ratio, emin 0.47 0.50
Max. void ratio, emax 0.78 0.74
Min. porosity, nmin 0.32 0.33
Max. porosity, nmax 0.44 0.43
Coefficient of uniformity, Cu 1.76 1.24
Coefficient of curvature, Cc 1.03 0.97
Specific gravity, Gs 2.65 2.65
USCS group symbol SP SP

Figure 1. Particle size distribution of binary soil mixtures

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 535

Test Setup and Procedure. Figure 2 shows an overview of the test setup used in this study
to measure thermal and electrical properties of the binary soil mixtures. A Miller box is
commonly used to measure the electrical resistivity of a soil sample, but the height of the Miller
box is too shallow to accommodate a thermal needle for thermal conductivity measurement.
Therefore, the authors customarily built a soil box with nominal dimensions of 110 mm (length)
x 50 mm (width) x 110 mm (height). Two 6.5 mm-thick stainless-steel electrode plates were
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

screwed onto the sides of the soil box. The soil box also had a 25 mm-diameter hole toward the
bottom of the front wall and the hole was connected to a water tank to saturate the soil sample.
The tests began by air-pluviating the binary soil mixtures from a hopper with a falling height
of 1400 mm and a hopper opening size of 5.1 mm to deposit a binary mixture sample into the
soil box. When the soil box was filled with the binary soil mixture sample, the pluviation was
stopped and the excess soil was trimmed away using a straight edge. After trimming, the weight
of the soil box filled with the binary mixture sample was measured, and the dry unit weight as
well as the void ratio of the deposited binary mixture sample were determined.
After depositing the binary mixture sample in the soil box, a 100-mm-long, 2-mm-diameter
thermal needle was slowly inserted into the center of the binary mixture sample and thermal
conductivity in a dry condition was measured using a portable thermal resistivity meter
(ThermTest TLS-100), followed by a careful removal of the thermal needle. The binary mixture
sample was then fully saturated with a 0.1% NaCl solution (i.e., 0.1% of NaCl and 99.9% of
deionized water by weight) by opening a water inlet valve of the soil box. The purpose of using
0.1% of NaCl solution, rather than 100% deionized freshwater, was to better observe the effect
of the percentage of finer particles on electrical resistivity of the binary mixture samples. Once
the water level reached the surface of the binary mixture sample, the water inlet valve was closed
and the weight of the wet mixture sample was measured to confirm that it was fully saturated.
The thermal needle was then inserted into the saturated binary mixture sample and thermal
conductivity in a saturated condition was measured.
After measuring thermal conductivity, an electrical resistivity test was performed on the
binary mixture samples saturated with the 0.1% NaCl solution. The two-electrode plates of the
soil box were connected to a Nilsson meter (Nilsson Soil Resistivity Meter Model 4000), and the
electrical resistance was read with the Nilsson meter and multiplied by a soil box factor to obtain
an electrical resistivity.
To verify the reproducibility of the test results, two mixture samples were prepared for each
percentage of finer particles and the thermal conductivity and electrical resistivity tests were
conducted on each sample. The values shown in the “Results and Discussions” section are
average values of the two measurements.

RESULTS AND DISCUSSIONS

Void Ratio of Deposited Binary Mixture Sample. As stated previously, the binary soil
mixtures were prepared at eight different mixing ratios based on weight. After depositing the
binary mixture sample in the soil box and trimming it, the void ratio (eM) of the binary mixture
sample was computed using Eq. (1):

𝐺𝑠 𝛾𝑤
𝑒𝑀 = 𝛾 − 1 (1)
𝑑

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 536

where gw = unit weight of water; gd = dry unit weight of soil = Ws/V where Ws = weight of the
deposited dry binary mixture sample; and V = volume of the soil box.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 2. A schematic of test setup

Figure 3 shows the void ratio eM of the binary mixture sample (average value of the two
samples) versus percentage of finer particles f. The void ratio eM decreased with increasing
weight percentage of the finer particles, reaching the smallest value (i.e., densest condition) at f 
30 %. When the finer particles were further added beyond 30%, eM increased as f increased. It
has been reported by many researchers that when finer particles are added to a matrix of coarser
particles the void ratio of the binary mixture initially decreases with an increase of finer particles
because the finer particles fill the voids between the coarser particles, achieving a minimum void
ratio of the mixture at an optimal amount of finer particles; further increase of finer particles
beyond the optimal fraction pushes coarser particles apart, leading to an increase of void ratio
(Chang et al. 2016; Choo et al. 2018; Kim and Seo 2019). The observation from this study is
consistent with these results.
Thermal Conductivity of Binary Mixture. ASTM D5334-14 (2014) specifies that the
diameter and height of a soil sample for thermal conductivity measurements should be greater
than 50 mm and 200 ± 30 mm, respectively. However, the height (= 110 mm) of the soil box

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 537

built in this study was smaller than the requirement. To test the effect of height of the soil sample
on the thermal conductivity, a series of thermal conductivity tests were conducted on
homogeneous, unsaturated F-55 sand with various sample heights, ranging from 100 to 200 mm,
in a cylindrical container. Results from these tests are shown in Figure 4. It shows that,
regardless of the sample height, the measured thermal conductivity values were very close to one
another, showing -5.2% to 2.6% difference when compared against the measured value with 200
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

mm sample height. Therefore, it was concluded that thermal conductivity measured with 110-
mm-height soil sample in this study was not significantly influenced by the shorter sample
height.

Figure 3. Void ratio of binary mixture sample versus percentage of finer particles

Figure 4. Thermal conductivities of F-55 sand for various sample heights

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 538

Thermal conductivity measurements of the binary mixtures in both dry and saturated
conditions are presented in Figure 5. The thermal conductivity of binary mixtures in saturated
conditions generally increased as the percentage of finer particles increased, peaked at f ranging
between 30 and 50%, and then decreased thereafter (this trend was less obvious for dry
conditions). Also, for the same percentage of finer particles, the thermal conductivity of binary
mixture samples in the saturated condition was 7.4 to 9.4 times (8.3 times in average) that in the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

dry condition. Roshankhah et al. (2021) conducted thermal conductivity tests on sand-silt
mixture samples, 72 mm in diameter and 76 mm in height, using a thermal probe. They reported
that the sand-silt mixture samples achieved a peak dry density and a peak thermal conductivity at
about 30-40% of silts in the mixture and the thermal conductivity of the saturated mixture was
more than seven times that of air-dry soil at the same density and effective stress. Experimental
results obtained from our study is in excellent agreements with these findings.

Figure 5. Thermal conductivity versus percentage of finer particles

Electrical Resistivity of Binary Mixtures. The electrical resistivities of the binary mixtures
were measured with the Nilsson meter using a 2-pin method, and the results are shown in Figure
6. Similar to thermal conductivity, electrical resistivity measurements increased as the
percentage of finer particles increased up to 30% and then decreased thereafter. Archie’s
empirical law (Archie 1942) suggests that electrical resistivity of water-saturated soils and rocks
increases with a decreasing porosity (n) with a power function relationship as follows:

𝐸𝑅𝑠𝑜𝑖𝑙
= 𝑛−𝑚 (2)
𝐸𝑅𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛

where ERsoil = electrical resistivity of saturated soil; ERsolution = electrical resistivity of pore
solution; n = porosity; and m = fitting parameter (termed as a cementation index by Archie).
Archie’s law was found to be valid for various porous materials, with exponents m ranging
approximately from 1.2 to 4.0. In this study, the value of m = 1.53 was obtained from a linear
regression analysis for the binary mixtures used in this study.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 539
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 6. Electrical resistivity in a saturated condition versus percentage of finer particles

SUMMARY AND CONCLUSIONS

The conclusions drawn from this study are summarized below:


• Void ratio eM of the binary soil mixtures decreased with increasing weight percentage of
the finer particles (f), reaching the smallest value (i.e., densest condition) at f  30 %.
When the finer particles were further added beyond 30%, eM increased as f increased.
• Thermal conductivity of the saturated binary mixture generally increased as f increased,
peaked at f  30-50%, and then decreased thereafter. Also, the thermal conductivity of
binary mixture sample in a saturated condition was 7.4 to 9.4 times (8.3 times in average)
that in a dry condition.
• Electrical resistivity of the binary mixture increased as f increased, peaked at f  30%,
and then decreased thereafter. Cementation index m in Archie’s empirical law was
determined to be 1.53 from a linear regression analysis for the binary mixtures used in
this study.

REFERENCES

Archie, G. E. (1942). “The electrical resistivity log as an aid in determining some reservoir
characteristics.” Trans. Am. Inst. Min. Metall. Pet. Eng. 146, 54–62.
ASTM. ASTM D5334-14. (2014). Standard Test Method for Determination of Thermal
Conductivity of Soil and Soft Rock by Thermal Needle Probe Procedure, ASTM
International, West Conshohocken, PA.
Brosseau, D., Kelton, J. W., Ray, D., Edgar, M., Chisman, K., and Emms, B. (2005). “Testing of
thermocline filler materials and molten-salt heat transfer fluids for thermal energy storage
systems in parabolic trough power plants.” J. Sol. Energy Eng. 127 (1): 109–116.
Cao, Z., Liang, X., Deng, Y., Wang, C., Wang, L., Zhu, R., and Zeng, J. (2021). “Influence of
multi-layered sediment characteristics on the thermal performance of buried submarine high-
voltage cables.” Ocean Engineering, 242, 110030.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 540

Chang, C. S., Wang, J.-Y., and Ge, L. (2016). “Maximum and minimum void ratios for sand–silt
mixtures.” Eng. Geol. 211, 7–18.
Choo, H., Lee, W., and Burns, S. E. (2018). “Estimating porosity and particle size for hydraulic
conductivity of binary mixed soils containing two different-sized silica particles.” J. Geotech.
Geoenviron. Eng., 144(1), 04017104.
De Vries, D. A. (1963). Thermal properties of soils. In: W.R. van Wijk, editor, Physics of plant
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

environment. North-Holland Publ. Co., Amsterdam. p. 210–235.


Friedman, S. P. (2005). “Soil properties influencing apparent electrical conductivity: a review.”
Computers and Electronics in Agriculture, 46, 45–70.
Kim, M., and Seo, H. (2019). “Evaluation of one- and two-parameter models for estimation of
void ratio of binary sand mixtures deposited by dry pluviation.” Granular Matter 21:71.
Laloui, L., Nuth, M., and Vulliet, L. (2006). “Experimental and numerical investigations of the
behaviour of a heat exchanger pile.” Int. J. Numer. Anal. Methods Geomech. 30 (8): 763–
781.
Roshankhah, S., Garcia, A. V., and Santamarina, J. C. (2021). Thermal Conductivity of Sand–
Silt Mixtures.” J. Geotech. Geoenviron. Eng., 2021, 147(2): 06020031.
Slegel, D. L., and Davis, L. R. (1977). “Transient Heat and Mass Transfer in Soils in the Vicinity
of Heated Porous Pipes.” J. Heat Transfer, 99(4): 541-546.
Sudha, K., Israil, M., Mittal, S., and Rai, J. (2009). “Soil characterization using electrical
resistivity tomography and geotechnical investigations.” J. of Applied Geophysics, 67(1): 74-
79.
Sun, Q., and Lu, C. (2019). “Semiempirical correlation between thermal conductivity and
electrical resistivity for silt and silty clay soils.” Geophysics 84(3): MR99-MR105.
Sreedeep, S., Reshma, A. C., and Singh, D. N. (2005). “Generalized relationship for determining
soil electrical resistivity from its thermal resistivity.” Experimental Thermal and Fluid
Science 29(2): 217–226.
Tang, A. M., Cui, Y. J., and Barnel, N. (2008). “Thermo-mechanical behavior of a compacted
swelling clay.” Géotechnique 58 (1): 45–54.
Tong, B., Gao, Z., Horton, R., Li, Y., and Wang, L. (2016). “An Empirical Model for Estimating
Soil Thermal Conductivity from Soil Water Content and Porosity.” Journal of
Hydrometeorology 17(2): 601-613.
Wallen, B. M., Smits, K. M., Sakaki, T., Howington, S. E., and Deepagoda, T. K. K. C. (2016).
“Thermal Conductivity of Binary Sand Mixtures Evaluated through Full Water Content
Range.” Soil Science Society of America Journal 80(3): 592-603.
Wang, J., Zhang, X., and Du, L. (2017). “A laboratory study of the correlation between the
thermal conductivity and electrical resistivity of soil.” Journal of Applied Geophysics 145:
12-16.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 541

Temperature Effects on Residual Shear Strength of Soil

Aidy Ung, S.M.ASCE1; Seyed Morteza Zeinali, S.M.ASCE2;


and Sherif L. Abdelaziz, Ph.D., A.M.ASCE3
1
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

VA. Email: unga@vt.edu


2
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
VA. Email: smzeinali@vt.edu
3
Charles E. Via, Jr. Dept. of Civil and Environmental Engineering, Virginia Tech, Blacksburg,
VA. Email: saziz@vt.edu

ABSTRACT

This study aims to assess temperature effects on the residual shear strength of soft clays.
Seasonal variation in temperature changes the engineering properties of soils. Temperature
fluctuation seasonally induces the destabilization of the ground near the surface and causes
geotechnical hazards such as thermally induced landslides. Therefore, in this study, a modified
temperature-controlled ring shear apparatus is used to change the temperature of clays and
evaluate the effect of the temperature on the residual shear strength. For this purpose, two clays
with two different mineralogies are selected and then residual shear strengths are measured at
room temperature (i.e., 20°C), 10°C, as well as at elevated temperatures of 30°C and 40°C. The
results for each of the considered clays at different temperatures are compared to assess the
temperature effects on the residual shear strength of soils. The results show a negligible impact
of temperature on the residual shear strength of Illite and Montmorillonite clay mixture
(Rhassoul clay). However, the residual shear strength of Kaolinite clay increases with either
heating or cooling. Although more investigation is required, the evolution of residual shear
strength with temperature is potentially dependent on the clay mineralogy.

INTRODUCTION

Climate change influences ground temperature (Beltrami and Kellman, 2003; Pollack et al.,
1998; Roy and Chapman, 2012), and the change in soil properties in response to climate change
leads to many geological and geomorphological hazards (McGuire and Maslin, 2012). The
seasonal variation of the ground temperature was found to weaken the shear strength along the
slip surface and trigger slow landslides movement (Shibasaki et al., 2016). As the climate change
is altering the thermal patterns around the globe, more natural and man-made slopes are
susceptible to instability (Davies et al., 2001; Wang et al., 2020). According to the U.S.
Geological Survey (2022), landslides are common geologic hazards that cost more than $1
billion in damage and between 25 and 50 death every year in the United States alone. The
residual shear strength is generally used for the design strength of soils with considerable
accumulation displacement, thus residual shear strength is an important parameter to investigate
landslides and reactivation potential (Skempton, 1964).
It is found that the increase in temperature increases the peak shear strength of the normally
consolidated clay (Hailemariam & Wutteke, 2022; Cekerevac & Laloui, 2004; Abuel-Naga et al.,
2006). However, the effect of the temperature of residual shear strength is not yet well
understood.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 542

The effect of temperature on residual shear strength provides a better understanding of the
mechanisms and thermal influence on occurrences of landslides. Understanding the effect of
temperature on the residual strength of soils could facilitate early risk mitigation associated with
landslides. Bucher (1975) found that the residual shear strength of two low-plasticity soils
(PI = 27, ϕr′ = 12.5° and PI = 30, ϕr′ = 25.6°) was not affected by the variability of temperature
from 10 to 60°C. However, the studies by Shibasaki et al. (2016, 2017) found that a drop in
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

temperature decreased the residual shear strength and friction angle of smectite-bearing soils,
resulting in many landslides occurring during winter independent of the increase in pore water
pressure. The study by Shibasaki et al. (2017) is limited to primarily smectite-rich soils and only
the effect of cooling process on the residual shear strength. This study aims to investigate both
the effect of cooling and heating on Kaolinite clay and Illite and Montmorillonite clay mixture
(Rhassoul clay) by evaluating the residual shear strength at room temperature, at 10 °C, as well
as at elevated temperatures of 30 and 40 °C.

SAMPLE PREPARATION AND METHODOLOGY

Soil Samples. Heating (30, 40 °C) and cooling (10 °C) tests were performed on two
commercial clays. A list of test samples and information on their corresponding index properties,
grain size distributions, and mineral assemblages are shown in Table 1. The index properties
were obtained by following ASTM D4318. The two selected clays are fine-grained soils; the
analysis of the grain size distribution was performed in accordance with ASTM D7928 on the
hydrometer test. X-ray powder diffraction (XRD) characterizations were conducted to identify
the mineralogy for both clays. A benchtop X-ray diffractometer (Rigaku Miniflex II), equipped
with a copper anode and 6-sample changer was used for mineral characterization. The specimens
used for the XRD tests were air-dried powder samples. Mineral identification was performed
using Match! Crystal Impact software package (Putz and Brandenburg, 2014).

Table 1. Testing Samples, Index Properties, Grain Size Distribution, and Mineral
Composition.

Kaolinite Clay Rhassoul Clay


PL (%) 32 43
LL (%) 67 138
PI 35 95
USCS Classification CH CH
Grain Size Clay (%) 60 84
Analysis Silt (%) 40 16
Clay Fraction % 60 84
Kaolinite (%) 96.7 0.1
Zeolite (%) 3.3 0
Illite (%) 0 29.4
Montmorillonite (%) 0 70.5

Ring Shear Apparatus and Set-Up. The residual shear strength is evaluated using a
Bromhead ring shear apparatus modified using JULABU Dyneo 1000F temperature controller
unit (Figure 1). The ring shear apparatus used in this experiment has the top platen modified as

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 543

suggested by Meehan et al. (2007) to eliminate the wall friction effect associated with top platen
intrusion. The setup of the ring shear device (Bromhead, 1979) was based on the guidance
provided by Castellanos and Brandon (2014).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Temperature-Controlled Modified Bromhead Ring Shear Apparatus.

The ring shear tests were performed following ASTM D6467. Reconstituted samples are
prepared at the liquid limit to minimize the air trapped when transferring the soil paste to the
sample container. Once the sample is placed in the apparatus, it is then submerged into di-
ionized water and consolidated at seven load increments to achieve normal stress between
approximately 7 and 300 kPa.
The soil samples are tested independently at different temperatures. For samples tested at
room temperature (20 °C), the sample is submerged and then maintained at room temperature.
The temperature is kept constant throughout the test and is monitored to ensure that the water in
the cell is 20°C. To change the temperature of the sample, the target temperature is achieved
using the temperature controller unit. Once the desired temperature is set, heated or cooled oil is
circulated in a coil placed inside the water bath, where the sample is seated.
In the ring shear experiments at different temperatures, the samples are first subjected to
consolidation stress of 7 kPa at room temperature. Once the consolidation under the first vertical
load is complete, the temperature is changed accordingly to 10, 30, and 40 °C and kept constant

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 544

during the test. A digital thermometer is used to monitor and ensure that throughout the test, the
temperature of water in the water bath does not fluctuate more than ± 0.5 °C from the intended-
testing temperature. In the consolidation process, each load is added after the previous load
reached primary consolidation. After the specimen was consolidated up to the final load of 300
kPa, the specimen was unloaded to the lowest stress to attain an overconsolidation state. The
specimen is then presheared at 7 kPa, corresponding to OCR of 41, under the rate of 15 degrees
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

per min (0.43 in/min) to form the failure plane. After a full rotation of the top platen, the axial
strain was monitored until the displacements were negligible. Then, the second load was applied
and at the end of the primary consolidation, the specimen was sheared at the rate of 0.024
degree/min or 0.0007in/min until it reaches a constant value. The shearing stage was then
stopped, and the specimen continued to be reloaded and sheared after reaching primary
consolidation.
The residual shear stresses corresponded to six normal stresses are then plotted, linearly fitted
with the intercept at the origin, to obtain a residual strength failure envelope and an estimated
residual friction angle. During the test, the water at respective temperatures is continuously
added to keep the sample saturated. Samples at elevated temperatures (30 and 40 °C) are
susceptible to quick evaporation of water in the water bath. To resolve this, a pipe is used to
deliver heated water pumped from a bowl placed on a hot plate to the water bath; the hot plate is
set at the testing temperature (Figure 2).

Figure 2. Pump Piping System to Keep Heated Samples Saturated.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 545

RESULTS AND DISCUSSION

The residual shear strength envelopes for the Kaolinite clay at different temperature is shown
in Figure 3. For Kaolinite clay, the residual shear strength increases slightly as the temperature
increases to 30 and 40 °C. The residual friction angle increases from 17.35° to 19.87° as the
temperature increases from 20 to 40 °C (Table 2). At the considered cooling temperature (10 °C),
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the residual shear strength also increases from 17.35° at 20 °C to 20.75° at 10 °C. This trend
suggests that there might be a minimum residual shear strength at a specific temperature, and the
plot of the residual shear strength against temperature follows a quadratic curve.

Figure 3. Residual Shear Strength Failure Envelope for Kaolinite Clay.

Table 2. Residual Friction Angles of Kaolinite Clay at Different Temperature.

T °C r (deg)
10 20.75
20 17.35
30 18.42
40 19.87

On the other hand, the residual shear strength envelope of Rhassoul Clay (Figure 4) remained
unchanged across the different considered temperatures. The residual friction angle remains the
same at about 9° at all considered temperatures (Table 3).
These results suggest that temperature does not impact the residual shear strength of Illite and
Montmorillonite clay mixture, but it does impact this for kaolinite clays.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 546
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4. Residual Shear Strength Failure Envelope for Rhassoul Clay.

Table 3. Residual Friction Angles of Rhassoul Clay at Different Temperature.

T °C r (deg)
10 9.73
20 9.23
30 9.70
40 9.62

It is also important to note that studies (Jefferson and Rogers, 1998; Shao et al., 2012;
Widjaja and Nirwanto, 2019; Yilmaz, 2011) have shown that temperature affects the liquid limit
of clay sample, and this effect depends on the mineralogy. Since all the samples in this study
were prepared at the liquid limit measured at room temperature, further investigation of the
relationship between temperature and liquid limits of each sample would allow for a better
understanding of the behavior of residual shear strength at different temperatures.

CONCLUSION

The effect of cooling and heating on two clays, one is a Kaolinite clay, and another is a
mixture of Illite and Montmorillonite clay minerals, was investigated by measuring the residual
shear strength at four different temperatures: 10, 20, 30, and 40 °C and developing failure
envelopes. A ring shear apparatus modified using a temperature controller unit was used to
measure the residual shear strength at different temperatures. It was found that the residual
friction angle increases when cooled and heated for the considered Kaolinite clay. This suggests
that there might be a minimum residual shear strength at a specific temperature. On the other

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 547

hand, the residual shear strength remains relatively unchanged for all temperatures tested for the
clay with Illite and Montmorillonite mixture (Rhassoul clay). This trend suggests that the effect
of temperature on the residual shear strength varies across clay mineralogy, thus soil samples
with various mineralogies need to be tested to better understand the effect of temperature on
residual shear strength. Since temperature also influences the liquid limit of clay, obtaining the
liquid limits of each sample at different temperatures would further shed light on this
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

relationship.

ACKNOWLEDGMENTS

This material is based upon work supported by the U. S. Army Research Laboratory and the
U. S. Army Research Office under contract numbers W911NF-16-1-0336, W911NF-17-1-0262,
W911NF-18-1-0068, and W911NF-20-1-0238. We also would like to acknowledge Dr. Jaradat
for performing XRD tests for this study. The discussions and conclusions presented in this work
reflect the opinions of the authors only.

REFERENCES

ASTM. ASTM Standard D4318 (2017-e1). Standard test methods for liquid limit, plastic limit
and shrinkage limit of soils. West Conshohocken, PA.
ASTM. ASTM Standard D6467 (2021-e1). Standard test method for torsional ring shear test to
determine drained residual shear strength of cohesive soils. West Conshohocken, PA.
ASTM. ASTM Standard D7928 (2021-e1). Standard test method for particle-size distribution
(gradation) of fine-grained soils using the sedimentation (hydrometer) analysis. West
Conshohocken, PA.
Abuel-Naga, H. M., Bergado, D. T., Ramana, G. V., Grino, L., Rujivipat, P., and Thet, Y.
(2006). Experimental evaluation of engineering behavior of soft Bangkok clay under elevated
temperature. Journal of geotechnical and geoenvironmental engineering, 132(7), 902-910.
Beltrami, H., and Kellman, L. (2003). “An examination of short-and long-term air–ground
temperature coupling.” Global Planet. Change, 38(3-4), 291-303.
Bromhead, E. (1979). “A simple ring shear apparatus.” Ground engineering, 12(5).
Bucher, F. (1975). Die Restscherfestigkeit natürlicher Böden, ihre Einflussgrössen und
Beziehungen als Ergebnis experimenteller Untersuchungen ETH Zurich].
Castellanos, B., and Brandon, T. (2014). Use and measurement of fully softened shear strength,
CGPR# 79. Center for Geotechnical Practice and Research, Blacksburg.
Cekerevac, C., and Laloui, L. (2004). Experimental study of thermal effects on the mechanical
behaviour of a clay. International journal for numerical and analytical methods in
geomechanics, 28(3), 209-228.
Davies, M. C. R., Hamza, O., and Harris, C. (2001). “The effect of rise in mean annual
temperature on the stability of rock slopes containing ice‐filled discontinuities.” Permafrost
Periglac., 12: 137-44.
Hailemariam, H., and Wuttke, F. (2022). An Experimental Study on the Effect of Temperature
on the Shear Strength Behavior of a Silty Clay Soil. Geotechnics, 2(1), 250-261.
Jefferson, I., and Rogers, C. D. F. (1998). “Liquid limit and the temperature sensitivity of
clays.”, Eng. Geol., 49(2), 95-109.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 548

McGuire, B., and Maslin, M. A. (2012). Climate forcing of geological hazards. John Wiley
Sons.
Meehan, C. L., Brandon, T. L., and Duncan, J. M. (2007). “Measuring drained residual strengths
in the Bromhead ring shear.” Geotech. Test. J., 30(6): 1–8.
Pollack, H. N., Huang, S., and Shen, P.-Y. (1998). “Climate change record in subsurface
temperatures: a global perspective.” Science, 282(5387), 279-281.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Putz, H., and Brandenburg, K. (2014). Match! -Phase Analysis using Powder Diffraction. Crystal
Impact GbR, Kreuzherrenstr, 102, 53227.
Roy, S., and Chapman, D. S. (2012). “Borehole temperatures and climate change: Ground
temperature change in south India over the past two centuries.”, J. Geophys. Res.-Atmos.,
117(D11).
Shao, Y. X., Shi, B., Liu, C., and Gao, L. (2012). “Experimental study on temperature effect on
engineering properties of clayey soils.” Adv. Mat. Res., vol. 512, pp. 1905-1918. Trans Tech
Publications Ltd, 2012.
Shibasaki, T., Matsuura, S., and Hasegawa, Y. (2017). Temperature‐dependent residual shear
strength characteristics of smectite‐bearing landslide soils. J. Geophys. Res-solid., 122(2),
1449-1469.
Shibasaki, T., Matsuura, S., and Okamoto, T. (2016). “Experimental evidence for shallow,
slow‐moving landslides activated by a decrease in ground temperature.” Geophys. Res. Lett.,
43(13), 6975-6984.
Skempton, A. (1964). “Long-term stability of clay slopes.” Geotechnique, 14(2), 77-102.
Wang, Q., Qi, J., Wu, H., Zeng, Y., Shui, W., Zeng, J., and Zhang, X. (2020). “Freeze-Thaw
cycle representation alters response of watershed hydrology to future climate change.”
Catena, 195: 104767.
Widjaja, B., and Nirwanto, A. F. (2019). “Effect of various temperatures to liquid limit, plastic
limit, and plasticity index of clays.” IOP conference series: materials science and
engineering.
Yilmaz, G. (2011). “The effects of temperature on the characteristics of kaolinite and bentonite.”
Sci. Res. Essays, 6(9), 1928-1939.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 549

Effects of Temperature on Volumetric Behavior of Soil Subjected to Freezing-Thawing


Cycles

Bohan Zhou1; Zihao Shang2; and Marcelo Sanchez3


1
Hydro China Huadong Engineering Cooperation, Hangzhou, Zhejiang, China; formerly, Dept.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

of Civil and Environmental Engineering, Texas A&M Univ., College Station, TX.
Email: zhou_bh2@hdec.com
2
Dept. of Civil and Environmental Engineering, Texas A&M Univ., College Station, TX.
Email: zihaoshang@tamu.edu
3
Dept. of Civil and Environmental Engineering, Texas A&M Univ., College Station, TX.
Email: msanchez@civil.tamu.edu

ABSTRACT

This paper presents an experimental investigation related to the effect of freezing-thawing (F-
T) cycles on the volumetric behavior of soils, with particular focus on the impact of the range of
freezing temperatures on soils’ response impact. A reconstituted specimen made up from a
commercial silt was subjected to F-T cycles in an 1D cell manufactured in a 3D-printer. The soil
specimen was fully saturated, and the test was conducted under open system conditions inside an
environmental chamber. Volume changes were recorded using a linear variable differential
transformer (LVDT) during cyclic F-T. The test results show that the impact of the range of the
freezing temperature is only relevant in the higher range of freezing temperatures (i.e., near the
freezing point), and it tends to reduce significantly as the minimum temperature reduces. The
study also shows that stress-history of the soil has a significant influence on the volumetric
behavior soils subjected to F-T cycles.

INTRODUCTION

Seasonal frozen soils dominate approximately 55% of the total earth's surface land (Zhang et
al., 2003). This area corresponds approximately to 55 million km2. In those regions, a good
understanding of the behavior of frozen soils is critical for a safe and economical design of new
civil infrastructure and for assessing the condition of existing structures. Furthermore, with the
discovery of fossil fuels in recent years, for example petroleum and gas hydrates near the Arctic
Circle as well as permafrost regions worldwide, frozen soils will likely become a topic of central
interest.
As a result of their wide distribution, frozen soils are of importance, and problems resulting
from frozen soils, especially soils subjected to cyclic freezing-thawing process are becoming
more significant constraints in many fields (i.e., engineering construction, waste disposal and
energy exploitation). This is because engineering properties of the soils such as strength,
stiffness, coefficient of permeability, and mechanical behavior change drastically with changes
in temperature and freezing-thawing cycles. These property changes give rise to many
engineering practical problems, bringing significant financial losses and safety problems, such as
distress of foundations due to thawing, leading to cracking of the super structure; railroad
distortion due to heaving of the soil; road surface damage due to thaw weakening etc. In
addition, the noticeable recent changes in climate worldwide, the increasing energy resource

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 550

exploitation demand in frozen ground, and the promising application of artificial ground freezing
technique for ground improvement, have enhanced the interest in the study of frozen soils.
It has been shown that, when the temperature falls below the water freezing point, not all the
pore-water present in the soil is converted into ice (Taber. 1930). There is an amount of unfrozen
water that attributed to capillary phenomena remains in liquid phase (see Figure 1). The amount
of unfrozen water content plays a critical role in evaluating the volumetric behavior of frozen
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

soil because it indicates how much water has been changed from water to ice (i.e., volume
expansion from water to ice is around 9%).
Different experimental techniques have been proposed to determine the amount of unfrozen
water in soils, e.g., calorimeter method (William, 1964), nuclear magnetic resonance (Tice et al.,
1982), differential scanning calorimetry (Kanitha and Reid, 2004), electrical conductivity (Mao
et al., 2018). In this paper, unfrozen water content was measured through electrical conductivity
method. For the soil we utilized in the research, the unfrozen water content changes as a result of
different freezing temperatures were measured at the beginning, because it is critical when
evaluating volumetric behavior of cyclic F-T tests under different F-T temperature ranges.

Figure 1. Schematic representation of frozen soil

The experimental study of the mechanical behavior of frozen soils has been generally based
on reconstituted specimens (e.g., Sayles 1974; Parameswaran 1980; Parameswaran and Jones
1981, Arenson et al. 2004; Chen et al. 2020). This is because gathering frozen undisturbed
samples from the field is challenging and expensive. In some works, undisturbed soil samples
were retrieved in-situ and were frozen in the lab before testing them (e.g., Cui et al. 2014; Tang
et al. 2018; Zhou et al. 2018). Only a few studies have been based on natural frozen samples
(e.g., Shastri et al., 2022). In terms of cyclic F-T tests, most of the experimental investigations
are based on studying the behavior of specimens that were already subjected to F-T cycles (either
in the lab or in the field) and comparing their responses against untreated soils. Konrad, 1989
reported that over-consolidated clays tend to expand when subjected to F-T cycles and that this
tendency is more notorious as the Over Consolidation Ratio (OCR) increases. Viklander, 1998
studied the permeability and volume changes of Normally Consolidated (NC) and Over
Consolidated (OC) soils subjected to F-T cycles. The permeability of NC samples decreases
because of the tendency of the soil to contract during F-T cycles. An opposite behavior was
observed in OC specimens. Both, OC and NC samples tend to a residual void ratio after a
number of cycles. Similar behavior was observed in silty soil studied by Qi et al., 2008 and Zhou

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 551

et al., 2017. Eigenbrod, 1996 found that fine-grained NC soils tend to contract during F-T cycles.
Eigenbrod et al., 1996 reported that no net volume changes were observed after F-T cycles in NC
samples with water content close to the plastic limit. Other studies related to soils subjected to F-
T cycles are focused on dynamic response, long-term cyclic loading, rheological properties,
resilient modulus, and influence of salt content and salt type.
However, the current knowledge on volumetric behavior of frozen soils subjected to F-T
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

cycles under different F-T temperature ranges with considerations of loading history is quite
limited. This paper introduced an experimental campaign focused on investigating such behavior
by performing a series of cyclic F-T tests in an environmental chamber. The unfrozen water
content under different freezing temperatures was also measured and a detailed discussion on
how F-T temperature range affected soil volumetric behavior was also presented.

TEST MATERIAL

Silt designated as SIL-CO-SIL 75 from the U.S. Silica Company was adopted for this
experiment. This silt was white in color and had a very low plasticity. Routine geotechnical tests
were conducted to determine the basic properties of this soil (Table 1). The liquid limit of the
sample was determined to be 25.3%. This material did not exhibit a plastic behavior, for example
the plastic limit could not be determined, therefore the soil was considered as non-plastic (NP).
The adopted water content was around 1.2 times the liquid limit and was set equal to 30%. The
specific gravity of this soil was 2.65. The initial saturation degree was 100% and initial void ratio
was 0.8.

Table 1. Summary of the soil properties

Property Value
Liquid Limit, LL (%) 25.30
Specific Gravity, Gs 2.65
Dry Density (kg/m3) 1350
Water content (%) 30
Initial Void Ratio, e0 0.80
Degree of Saturation, Sr (%) 100

EXPERIMENTAL SETUP

A high-quality 3D printer was used to manufacture the one-dimensional cell adopted in this
research. It consists of the following main components: a base pedestal; an inner-ring (to house
the soil); a plunger (to apply the vertical stress); an outer-ring (to provide a seal all around and to
support the plunger); a LVDT (linear variable differential transformer); and a mount (which
assembles all the components together). Figure 2a_ and 2b show the adopted setup and the setup
inside the chamber. O-Ring seals between outer-ring and pedestal were included to ensure a
watertight setup. The pedestal was designed with drainage at the bottom connected to a source of
free water to produce an open system and allow free drainage. The inner ring can host soil
specimens 39mm height and 39mm diameter. Porous stones were placed at the bottom and top of
the samples. The vertical load was applied as a dead load through the plunger.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 552
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

a) b)

Figure 2. a) 1-D cell adopted in the experiments b) 1-D cells inside the CSZ chamber.

The LVDT (SE-750-500) attached to the top of the plunger tracked the vertical
displacements during the test. This LVDT is specially designed to operate under freezing
temperature (-20°C to 70°C). The F-T cycles were conducted inside a CSZ 0.45m3 (16ft3)
environmental chamber capable of temperature (-40°C to 179°C) and relative humidity (5% to
98%) computer-controlled cycling (Figure 3a)). The temperature inside chamber was also
measured using an built in thermocouple. The set-up was calibrated for the range of temperature
contemplated in this research.
The oven-dried silt was mixed with distilled water at the target water content (i.e. ~w=30%)
and the mixtures were left inside hermetic plastic bags for 24 hours to attain homogenized
specimens. All the samples were prepared inside the cell in slurry state. A light tapping was
applied to remove any possible air-bubble. LandMapper ERM-02 (Figure 3b) was selected in this
preliminary research to measure the electrical conductivity (EC) of the silty soil. It is a four-
electrode portable device, easy to calibrate and use.

(a) b)

Figure 3. Test equipment a) Environmental chamber; b) LandMapper ERM-02

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 553

UNFROZEN WATER CONTENT MEASUREMENTS

To estimate the amount of the unfrozen water retained at different freezing temperature
electrical conductivity method was used. The LandMapper ERM-02 device was adopted in this
study (see Figure 3b) following the method and procedure reported in Mao et al., (2018. A large
amount of silt was mixed with distilled water to reach a water content of 30% (slurry sample) in
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

a large glass beaker as discussed before, and the beaker was then sealed with a plastic bag for a
24 h settlement. Liquid on top of the sediment will be used as the interstitial liquid for this
method. Thermometer probes and the LandMapper ERM-02 were placed in the liquid and then
glass beaker with liquid was placed in the environmental chamber. Reading of temperatures as
well as electrical conductivity were taken outside of the chamber by thermometer and
LandMapper ERM-02 separately. Chamber temperature was controlled from positive to negative
to get a curve relating EC and temperature values of the interstitial liquid alone. Additional soil
sample with water content of 30% was prepared and move it into the mold as discussed before.
Both the thermometer probe and LandMapper ERM-02 probe were inserted into the middle of
the soil sample, and then whole set-up was placed into the environmental chamber. Reading of
temperatures as well as electrical conductivity were taken outside of the chamber by
thermometer and LandMapper ERM-02 separately. Chamber temperature was controlled from
positive to negative to get a curve relating EC and temperature values of the soil sample.
The saturation degree of unfrozen water, Sluw, was estimated based on Archie’s law: 𝑆𝑙𝑢𝑤 =
𝐸𝐶𝑠 ×𝑛−𝑝 ×𝑆𝑙 −𝑞
; where ECs and ECw are the electrical conductivities of the frozen soil and
𝐸𝐶𝑤
interstitial liquid, respectively, n is the porosity, p is an exponent related to soil structure, and q is
an exponent associated with S1 (Sl=1 in this case). The following values have been suggested for
the parameters involved in the equation above: q~2.0 and 1.4<p< 2.0.
The EC of soil was shown in Figure 4a), the freezing point of the interstitial liquid was
around 0℃ and the EC of frozen soil decreased sharply at the very beginning (i.e., 0 to -2℃)
when temperature decreases negatively, same thing happened with Sluw. But, for temperature
below -2℃, Sluw remains almost stable and very low (see Figure 4b)). This appears to be because
for this soil, and the existing experimental conditions, virtually all the water was frozen after the
0 to -2℃ range freezing process.

Figure 4. EC and Unfrozen water retention curve

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 554

CYCLIC TESTS UNDER DIFFERENT F-T TEMPERATURE RANGE

Six tests in three groups were conducted to investigate the potential effects of the temperature
range on soils volume change subjected to F-T cycles. Temperature variations between -0.3°C
and 5°C were applied to Normally Consolidated (NC) specimens under 'v=10kPa and Over
Consolidated (OC) samples under the same 'v but previously loaded up to 'v=500kPa. A
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

similar protocol was followed to the other two groups of specimens subjected to temperature
changes between [-7°C to 5°C] and [-20°C to 5°C]. In all the cases a maximum of 8 F-T cycles
were investigated. In all the tests the volume change was stable after the 6 cycles. The primary
results are presented in Figure 5. It can be observed that for the two states considered here (i.e.,
NC and OC), NC soil tended to contract while OC soil tended to expand. The contraction amount
for these three NC groups was almost the same, but for OC soil the range [-0.3°C to 5°C]
induced less final expansion than the other F-T temperature two ranges. Furthermore, the
sequence [-0.3°C to 5°C] induced less volume changes during F-T cycles than the other two
temperature ranges, also there were practically no differences between the series [-7°C to 5°C] and
[-20°C and 5°C]. According to unfrozen water content measurements, this behavior could be
anticipated because unfrozen water content was almost the same for sample under -7°C and -20°C.

a)

b)

Figure 5. Effect of temperature range during F-T cycles on soil behavior a) NC soil; b) OC
soil

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 555

DISCUSSION AND CONCLUSION

Through this experimental study, a better understanding of the volumetric behavior of frozen
soils subjected to different loading histories, and different temperature ranges during F-T cycles
was achieved. As for the effect of the temperature range on soils subjected to F-T cycles, a
marked difference in terms of volumetric behavior was observed when comparing the range [-
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

0.3℃ to 5℃] and [-7℃ to 5℃]. However, almost no difference was observed in the volumetric
response of the samples when comparing the ranges [-7℃ to 5℃] and [-20℃ to 5℃]. This
behavior can be explained when inspecting the unfrozen-water content retention curve. It is
observed that at the very beginning of the cooling stage (i.e., close to freezing point) the amount
of unfrozen water decreased sharply, but at lower temperatures the Sluw remained almost stable
and very low. It was observed that for this particular soil, almost all the water is in the ice form
when the temperature is below -2℃. This behavior explained why little or no noticeable
differences in terms of volume changes are observed when comparing the response of sample
during F-T cycles at the low temperature ranges (i.e., [-7℃ to 5℃] against [-20℃ to 5℃]).
The financial support from the United State National Science Foundation, award number:
2034204 is acknowledged.

REFERENCES

Arenson, L. U., Johansen, M. M., and Springman, S. M. (2004). Effects of volumetric ice content
and strain rate on shear strength under triaxial conditions for frozen soil samples. Permafrost
Periglac Process. 15, 261–271.
Chen, H., Guo, H., Yuan, X., Chen, Y., and Sun, C. (2020). Effect of Temperature on the
Strength Characteristics of Unsaturated Silty Clay in Seasonal Frozen Region. KSCE J Civ
Eng. 24, 2610–2620.
Cui, Z. D., He, P. P., and Yang, W. H. (2014). Mechanical properties of a silty clay subjected to
freezing-thawing. Cold Reg Sci Technol. 98, 26–34.
Eigenbrod, K. (1996). Effects of cyclic freezing and thawing on volume changes and
permeabilities of soft fine-gained soils. Canada Geotechnical Journal, 33, 529-537.
Eigenbrod, K., Knutsson, S., and Sheng, D. (1996). Pore-Water Pressures in Freezing and
Thawing Fine-Grained Soils. Journal of Cold Regions Engineering, 10, 77-92.
Kanitha, T., and Reid, D. S. (2004). DSC and NMR relaxation studies of starch water
interactions during gelatinization. Carbohydrate Polymers, 58(3), 345–358.
Konrad, J. (1989). Effect of freeze-thaw cycles on the freezing characteristics of a clayey silt at
various overconsolidation ratios. Canada Geotechnical Journal, 26, 217-226.
Mao, Y., Romero, E., and Gens, A. (2018). Ice formation in unsaturated frozen soils. 7th
International Conference on Unsaturated Soils, Hong Kong, China.
Parameswaran, V. R. (1980). Deformation behaviour and strength of frozen sand. Can Geotech
J. 17, 74–88.
Parameswaran, V. R., and Jones, S. J. (1981). Triaxial testing of frozen sand. J Glaciol. 27, 147–
55.
Qi, J., Ma, W., and Song, C. (2008). Influence of freeze-thaw on engineering properties of a silty
soil. Cold Regions Science and Technology, 53(3), 397-404.
Sayles, F. H. (1974). Triaxial Constant Strain Rate Tests and Triaxial Creep Tests on Frozen
Ottawa Sand. US Army Corps Eng Cold Reg Res Eng Lab Tech Rep.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 556

Shastri, A., Sánchez, M., Gai, X., Lee, M. Y., and Dewers, T. (2021). Mechanical behavior of
frozen soils: Experimental investigation and numerical modeling. Comput. Geotech. 138,
104361.
Taber, S. (1930). The mechanics of frost heaving. The Journal of Geology, 38(4), 303-317.
Tang, L., Cong, S., Geng, L., Ling, X., and Gan, F. (2018). The effect of freeze-thaw cycling on
the mechanical properties of expansive soils. Cold Reg Sci Technol. 145, 197–207.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Tice, A. R., Oliphant, J. L., Nakano, Y., and Jenkins, T. F. (1982). Relationship between the ice
and unfrozen water phases in frozen soil as determined by pulsed nuclear magnetic resonance
and physical desorption data. Cold Regions Research and Engineering Lab Report, 82-15.
Viklander, P. (1998). Permeability and volume changes in till due to cyclic freeze-thaw. Canada
Geotechnical Journal, 35(3), 471-477.
Williams, P. J. (1964). Unfrozen water content of frozen soils and soil moisture suction.
Géotechnique, 14(3), 231-246.
Zhang, T., Barry, R. G., Knowles, K., Ling, F., and Armstrong, R. L. (2003). Distribution of
seasonally and perennially frozen ground in the Northern Hemisphere. Conference:
Proceedings of the 8th International Conference on Permafrost At: Zurich, Switzerland, 2.
Zhou, B., Sanchez, M., Shastri, A., and Lee, J. (2017). Mechanical Behavior of Soil Subjected to
Freezing-Thawing Cycles. Second Pan-American Conference on Unsaturated Soils,
November 12–15, 2017, Dallas, Texas.
Zhou, Z., Ma, W., Zhang, S., Mu, Y., and Li, G. (2018). Effect of freeze-thaw cycles in
mechanical behaviors of frozen loess. Cold Reg Sci Technol. 146, 9–18.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 557

Particle Shape Effects in 3D DEM Simulations of Angle of Repose

C. S. Sandeep, A.M.ASCE1; and T. M. Evans, M.ASCE2


1
Postdoctoral Researcher, School of Civil and Construction Engineering, Oregon State Univ.,
Corvallis, OR. Email: chittas@oregonstate.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Professor, School of Civil and Construction Engineering, Oregon State Univ., Corvallis, OR.
Email: matt.evans@oregonstate.edu

ABSTRACT

Crushed rock is a common granular material in civil engineering, mining, and railway
applications. Numerous prior studies have shown that design-scale mechanical behavior strongly
depends on the morphological features of the particles. In this study an efficient technique to
import realistic particle geometries into DEM simulations of bulk granular materials is provided.
First, a suite of 300 synthetic 3D particle geometries (referred to as particle avatars) of varying
sizes and shapes is prepared using a computational physics engine. Second, relationships
between widely used 2D and 3D shape descriptors are obtained for the particle avatars. From the
2D-3D shape relationships it is observed that the true sphericity (3D) is well correlated with
perimeter sphericity (2D). Therefore, some 3D shape features can be reasonably estimated from
2D digital images of grains. Finally, DEM simulations are carried out to investigate angle of
repose for several clump geometries. The bubble pack technique was used to create pebble
clumps with varied numbers of pebbles. Results show that the repose angle is more strongly
dependent on the shape of the created clump (collection of pebbles) than by the shape features of
the original particle geometry. This research shows that the shape features of the clumps used in
simulations affect the results, regardless of the procedures used to get 3D geometries of particles.

INTRODUCTION

Granular materials are the second most processed material after water, and have received
extensive research. However, due to their multiphase properties, some issues such as fluid-solid
transition and flowability of granular materials remain challenging (Van Der Meer, 2017; Ma et
al., 2019). Repose angle is a quick test for assessing the flowability as it correlates with the
strength of particle-particle interactions and is also used to quantify strength of granular
materials. Several researchers have investigated the role of factors including moisture, gradation,
friction coefficient, aspect ratio, and material properties on the angle of repose (e.g., Chen et al.
2020; Ma et al. 2020). However, the influence of particle morphology on angle of repose is
relatively less studied, though it is known to play an important role in the engineering behavior
(e.g., Santamarina and Cho 2004, Cho et al. 2006).
The relationship between particle morphology and macroscale behavior, including angle of
repose, can be readily explored with the discrete element method (DEM) by using real 3D
particle shapes as model inputs. To this end, several techniques are available to characterize 3D
particle shape, including X-ray microcomputed tomography (µCT), laser scanning, and
photogrammetry, but while highly accurate, they are also expensive, time-consuming, and not
routinely available for most practical applications. These approaches are also particle-size-
dependent – small particles (sand grains) are well-handled by µCT while larger particles

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 558

(crushed rock) require a different approach. For large particle sizes in the field, it is difficult to
completely capture the 3D particle geometry using photogrammetry or laser scanning. These
methods can be considered to give 2.5D particle shape.
In order to include realistic particle geometries in numerical simulations, an effective
approach is required. One simple method is to assess the possibility of obtaining 3D shape
characteristics from 2D digital images of grains. For a few granular materials, researchers (e.g.,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Rorato et al. 2019) studied correlations between 2D and 3D shape descriptors. However, these
correlations should be preferably independent of particle orientation, material type, and size for
practical applications such as modeling of large particle sizes in the mining sector.
In this study, an effective technique to export realistic particle geometries into numerical
simulations is presented by analyzing relationships between 2D and 3D shape descriptors for
realistic 3D particle geometries created via a computational physics engine. Thereafter, the role
of particle geometry and clump morphology on the flowability of granular materials is
investigated through carrying out angle of repose numerical experiments via discrete element
method (DEM).

METHODOLOGY

Brief Introduction to DEM. DEM is a numerical tool for simulating behavior of granular
assemblies (Cundall & Strack, 1979). DEM uses Newton's second law to explicitly account for
inter-particle interactions and track particle trajectories. This method made significant
contributions in understanding the effect of particle shape on mechanical response of sands and
other granular materials (e.g., Zhang et al. 2020). For more information regarding this method
and the effectiveness of DEM in various applications readers can refer to O’Sullivan (2011). In
the current study, numerical experiments are carried out using the commercially available three-
dimensional DEM code PFC3D (Itasca Consulting Group, 2022).
Irregular Particle Shapes in DEM. Two techniques are widely employed to simulate the
influence of particle shape in DEM. To account for the resistance to rotation generated by
particle shape, the first way is to use a rheology-type rolling resistance model to define the
constitutive relationship at the inter-particle contact and a more direct method is to import
irregular particle morphologies directly into DEM simulations.
Blender (Blender Online Community 2022), an open-source 3D computer graphics program
is used in the current work to produce mesh geometries of synthetic particles referred to as
‘avatars’. Sandeep et al. (2022) demonstrated that this technique can produce realistic particle
avatars of various shapes, which can represent shape features of commonly used grains in
geotechnical engineering. The particle avatars' surface files are imported into the DEM
simulation, and clumps are generated through the multi-sphere approach proposed by Taghavi
(2011). The degree to which the generated clump closely resembles the exact shape of the
irregular particle is determined by two parameters, α and β. The ratio of the smallest to the
largest pebble within the clump is defined as the value of α, and β refers to the angular measure
of smoothness. Figure 1 depicts representative clumps formed by varying the values of α and β
for a given 3D particle geometry.
Shape Characterization. For characterizing the shape of the 3D particle geometry (from
‘avatar’) and generated clumps (as shown in Figure 1), commonly used 2D and 3D shape
descriptors in the literature are employed. True sphericity and roundness in 3D, and perimeter
sphericity and aspect ratio in 2D, are some popular shape descriptors. Wadell (1932) defined true

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 559

sphericity (3D) as the ratio of the surface area of a sphere with the same volume as the particle to
the surface area of the particle itself. Roundness (3D) is calculated as the proportion of the
average radius of curvature of a particle's corners to the radius of the largest inscribed circle.
Perimeter sphericity (2D) is defined as the ratio of the perimeter of a circle with the same area as
the particle to the particle's actual perimeter. Aspect ratio (2D) is defined as the ratio of major
axis length to intermediate axis length.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1. Particle surface and generated clumps with varying numbers of pebbles

Simulation Setup. In this study, strength of granular materials is evaluated through the
repose angle test. Figure 2 shows repose angle test setup used in this work. The container (3 cm
in diameter, 12 cm in height) is filled with at least 4000 particles with size range of 2-6 mm
(uniformly distributed PSD) and density of 2.65x106 kg/m3. Parameters used in the simulation
are given in Table 1. For the sake of computational efficiency, a linear-spring contact model was
used in this study, which can produce similar microscopic and macroscopic responses to the
Hertz–Mindlin model (Zhao et al. 2018a). The normal and tangential contact stiffnesses were set
to 5.0x106 N/m. The particles are generated and filled in the container at random and the inter-
particle friction is set to zero during the initial filling process. The particles are then allowed to
settle under gravity. During the testing, the cylinder is gradually lifted at a velocity of 50 mm/s,
and the particles are collected in the bottom plate (3.5 mm in height, 6 cm in diameter).
Automatic time-step scaling is enabled during particle simulation to reduce computational time.

Table 1. Parameters used in the simulation

Parameters Value
Damping ratio 0.3
Normal stiffness 5.0 x 106 N/m
6
Tangential stiffness 5.0 x 10 N/m
Inter-particle friction 0.5
Particle-cylinder friction 0.0
Particle-base plate friction 1.0

RESULTS AND DISCUSSION

Relationship Between 2D and 3D Shape Descriptors. Sandeep et al. (2022) generated a


database of 300 virtual particle avatars. They obtained 2D and 3D shape characteristics for the

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 560

particle avatars that could represent shape features of various geotechnical grains. Figure 3
shows relationship between perimeter sphericity (2D) and true sphericity (3D). The values of
perimeter sphericity are based on the mean from 100 random projections. For calculating surface
area and consequently true sphericity, the particle's surface mesh is triangulated. Figure 3 shows
that as the true sphericity decreases, so does the perimeter sphericity. Lower perimeter sphericity
and true sphericity values correspond to irregular particles. For example, the particle avatar
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(surface) shown in Figure 1 has a true sphericity value of 0.88, whereas a sphere has a true
sphericity value of one. According to Figure 3, true sphericity values could range between 0.75
and 0.92 for a range of perimeter sphericity values between 0.9 and 0.95 (representative for
Peace River sand). The estimated range of true sphericity values from Figure 3 matches well
with the true sphericity values (mean: 0.824, standard deviation: 0.091) determined via micro-CT
for Peace River sand (after Li et al. 2022). Therefore, some 3D shape characteristics like true
sphericity can be approximated from simple 2D shape characteristics like perimeter sphericity.

Figure 2. Repose angle test setup

Effect of Number of Pebbles on Shape. Clumps are generated for a particle avatar through
the multi-sphere approach (as shown in Figure 1) and each sphere is referred to as pebble. The
generated shape of the clump gets closer to the actual shape of the particle geometry as the
number of pebbles increases. Figure 4 shows effect of number of pebbles considered to generate
a clump versus 3D shape characteristics. The values of true sphericity and 3D roundness for the
original particle avatar are 0.88 and 0.53, respectively. From Figure 4, the true sphericity and
roundness values decrease sharply as the number of pebbles increases from 1 to 6, and they
nearly reach a plateau after about 46 pebbles. However, there is no consistent relationship
between the number of pebbles and 3D shape descriptor (such as true sphericity). This is due to
differences in the rate of change in volume and surface area as the number of pebbles increases.
This analysis shows that clumps of various shape characteristics can be generated for any particle
geometry.
Effect of Particle Shape on Angle of Repose. In the current study, DEM simulations of
angle of repose experiments are carried out for the clumps that are created with 1, 2, 6, 11, and
46 pebbles. This is to limit the computational time as clumps generated through larger number of
pebbles require greater computational time. The angle of repose is estimated by considering the
tangent of the height of the granular pile to its base radius. Figure 5 shows variation in angle of

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 561

repose with number of pebbles used to create the clump. The number of pebbles corresponding
to 1 refers to a sphere as shown in Figure 1. According to Figure 5, the angle of repose increased
with the number of pebbles and nearly reached a plateau. The angle of repose values for clumps
generated from 1, 2, 6, 11, and 46 pebbles are 22.7°, 30.8°, 35.5°, 34.3°, and 36.6°, respectively.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Variation in true sphericity (3D) versus perimeter sphericity (2D)

Figure 4. Variation in true sphericity and roundness with the number of pebbles used to
generate clumps

Figure 6a shows variation in angle of repose with aspect ratio. For the aspect ratio range
considered in this study, no definite relationship between repose angle and aspect ratio is
observed. Zhao et al. (2018b), on the other hand, demonstrated that the value of angle of repose
increases with increasing aspect ratio (range 1-5) for rod-shaped particles. Variation of angle of
repose with regularity is shown in Figure 6b. Regularity is defined as the average value of true

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 562

sphericity and 3D roundness. Figure 6b shows that the angle of repose values of the pile
increases with decreasing regularity. Figures 5, 6a, and 6b demonstrate that changing the number
of pebbles used to generate the clump for a given particle geometry (true sphericity: 0.88;
roundness: 0.53) affects shape of the generated clump, which in turn affects the flowability of
granular materials. Therefore, regardless of the initial particle geometry (e.g., surface in Figure
1), the shape of generated clumps influences granular material flowability.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5. Variation in angle of repose versus with the number of pebbles used to generate
clumps

Figure 6. Angle of repose versus (a) aspect ratio of the clump (b) regularity of the clump

CONCLUSIONS

This study demonstrated an efficient method for importing realistic particle geometries into
discrete element method (DEM) simulations of granular media. Furthermore, the effect of clump
morphology on flowability is investigated using numerical repose angle experiments. The main
findings can be summarized as:

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 563

• True sphericity (3D) can be approximated from perimeter sphericity (2D) of randomly
oriented particles in a 2D digital image.
• Using a multi-sphere technique, clumps with a wide range of 3D shape features can be
created for a given particle geometry.
• Repose angle increases with decreasing regularity of the clumps. Therefore, flowability
of the granular material is determined by the shape features of the clumps, irrespective of
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

the approach employed to obtain 3D particle geometry.

ACKNOWLEDGEMENTS

This material is based upon work supported by the Oregon Metals Initiative and ESCO
Group. This support is gratefully acknowledged. The work has benefitted from many
conversations with Abram Hernandez and Kevin Stangeland.

REFERENCES

Blender Online Community. (2021). Blender - a 3D modelling and rendering package. Blender
Foundation, Blender Institute, Amsterdam.
Chen, H., Zhao, S., and Zhou, X. (2020). “DEM investigation of angle of repose for super-
ellipsoidal particles.” Particuology, 50, 53-66.
Cho, G.-C., Dodds, J., and Santamarina, J. C. (2006). “Particle Shape Effects on Packing
Density, Stiffness, and Strength: Natural and Crushed Sands.” J. Geotech. Geoenviron. Eng.,
132 (5): 591–602.
Cundall, P. A., and Strack, O. D. (1979). “A discrete numerical model for granular assemblies.”
Géotechnique, 29(1), 47-65.
Itasca Consulting Group. (2022). PFC 6.0 manual. Minneapolis, MN.
Li, L., Sun, Q., and Iskander, M. (2022). “Efficacy of 3D dynamic image analysis for
characterising the morphology of natural sands.” Géotechnique.
Ma, Y., Evans, T. M., Philips, N., and Cunningham, N. (2019). “Modeling the effect of moisture
on the flowability of a granular material.” Meccanica, 54(4), 667-681.
Ma, Y., Evans, T. M., Philips, N., and Cunningham, N. (2020). “Numerical simulation of the
effect of fine fraction on the flowability of powders in additive manufacturing.” Powder
Technology, 360, 608-621.
O’Sullivan, C. (2011). Particulate discrete element modelling: a geomechanics perspective.
CRC Press.
Rorato, R., Arroyo, M., Andò, E., and Gens, A. (2019). “Sphericity measures of sand grains.”
Engineering Geology, 254: 43–53.
Sandeep, C. S., Hernandez, A., Stangeland, K., and Evans, T. M. (2022). Shape characteristics
of granular materials through realistic particle avatars. (Under Review).
Santamarina, J. C., and Cho, G. C. (2004). “Soil behaviour: The role of particle shape.” The
Skempton conference: Proceedings of a three-day conference on advances in geotechnical
engineering, organised by the Institution of Civil Engineers and held at the Royal
Geographical Society, London, UK, on 29–31 March 2004, 604-617.
Taghavi, R. (2011). “Automatic clump generation based on mid-surface.” In Proceedings, 2nd
international FLAC/DEM symposium, Melbourne, 791-797.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 564

Van Der Meer, D. (2017). “Impact on granular beds.” Annual review of fluid mechanics, 49, 463-
484.
Wadell, H. (1932). “Volume, shape, and roundness of rock particles.” The Journal of Geology,
University of Chicago Press, 40(5), 443–451.
Zhang, N., Evans, T. M., Zhao, S., Du, Y., and Zhang, L. (2020). “Discrete element method
simulations of offshore plate anchor keying behavior in granular soils.” Marine
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Georesources & Geotechnology, 38(6), 716-729.


Zhao, H., An, X., Gou, D., Zhao, B., and Yang, R. (2018b). “Attenuation of pressure dips
underneath piles of spherocylinders.” Soft Matter, 14(21), 4404-4410.
Zhao, S., Evans, T. M., and Zhou, X. (2018a). “Effects of curvature-related DEM contact model
on the macro-and micro-mechanical behaviours of granular soils.” Géotechnique, 68(12),
1085-1098.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 565

Internal Structure and Breakage Behavior of Biogenic Carbonate Sand Grains

Elieh Mohtashami1; C. Guney Olgun, Ph.D.2; Chenglin Wu, Ph.D.3; and Tara Selly, Ph.D.4
1
Ph.D. Candidate, Dept. of Civil, Architectural, and Environmental Engineering, Missouri Univ.
of Science and Technology, Rolla, MO. Email: mohtashamie@mst.edu
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

2
Assistant Professor, Dept. of Civil, Architectural, and Environmental Engineering, Missouri
Univ. of Science and Technology, Rolla, MO. Email: olgun@mst.edu
3
Assistant Professor, Dept. of Civil, Architectural, and Environmental Engineering, Missouri
Univ. of Science and Technology, Rolla, MO. Email: wuch@mst.edu
4
Assistant Director, X-Ray Microanalysis Core Facility, Office of Research, and Research
Assistant Professor, Dept. of Geological Sciences, Univ. of Missouri, Columbia, MO.
Email: sellyt@missouri.edu

ABSTRACT

This study investigates the mechanical behavior of biogenic carbonate sands from Puerto
Rico at grain-scale level. Micro-computed tomography has also been used to get insights on the
internal structure of these particles before and after loading. The crushing strength of these
particles are smaller comparing to the values reported for silica sands. It has also been shown
that these particles have complex internal structure including a network of pores connected with
channels. This study also demonstrates the effect of intragrain structure of biogenic carbonate
sands and shows how internal grain structure plays a role on particle fracture.

INTRODUCTION

Carbonate sands can be found in offshore and tropical environments. They account for about
40% of ocean floors and are globally distributed, including along the coasts of Australia, Persian
Gulf, Gulf of Mexico, South of China, and Ireland (Airey et al. 1988, Coop 1988, Hassanlourad
et al. 2008, Salem et al. 2013, Xiao et al. 2017). Across the United States, they can be found in
Hawaii, coasts of Florida and Puerto Rico (Sandoval and Pando 2012, Morales-velez et al. 2015).
Unlike silica sands, that are alluvial sediments made of quartz, carbonate sands are formed as a
result of the deposition of skeletal remains of marine organisms (biogenic) or by the chemical
precipitation of calcium carbonate in a calcium rich environment (non-biogenic) and contain
high amount of calcium carbonate (Fookes 1988).
Differences have been reported in the macro-scale behavior of carbonate and silica sands
including their compression behavior, shear behavior and dynamic behavior (Poulos et al. 1982,
Golightly and Hyde 1988, Coop 1990, Hyodo et al. 1996, Kwag et al. 1999, McDowell and
Harireche 2002, Porcino et al. 2008, Brandes 2011, Beemer et al. 2019a, Saeidaskari et al. 2020,
Rasouli et al. 2021). This can be attributed to the unique characteristics of the carbonate sand
particles as the result of their deposition environment. While silica sand grains are bulky, round
and practically solid with miniscule amount of pores, carbonate sand particles are usually
elongated and angular with high internal porosity (Cil and Alshibli 2012, Al Mahbub and Haque
2016, Ma et al. 2019, Li et al. 2020, Mohtashami et al. 2022).
Although more studies have become available recently considering the macro-scale behavior
of carbonate sands, a few studies have been focused on the mechanical behavior of carbonate

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 566

sands at grain-scale level (Orense et al. 2013, Beemer et al. 2019b, Ma et al. 2019, Zhang et al.
2019, Li et al. 2020, Lv et al. 2020, Xuehui et al. 2020, Kuang et al. 2021, He et al. 2022). There
is still a need for understanding the behavior of carbonate sands particularly considering the
internal structure of these particles.
With the development of the X-ray micro-computed tomography (micro-CT) technique,
insights are provided into the 3D internal structure of materials. Particularly, in the field of
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

geotechnical engineering, this technique has become a popular tool in the studies at grain-scale
levels. X-ray CT is a non-destructive imaging technique that produces high resolution 3D image
by measuring the level of attenuation of X-ray intensity, for a material bombarded with X-rays is
measured. The final product is a series of gray-scale slice images. The level of the X-rays
attenuation is a function of the specimen chemical composition, its geometry, and the energy
level of the X-ray source. This technique requires minimal sample preparation and has relatively
quick scanning time (Druckrey and Alshibli 2016, Zhao et al. 2020).
This study took an experimental approach to investigate the mechanical behavior of
carbonate sand particles using uniaxial compression tests and the internal structure of the
particles using micro-computed tomography. The effect of the internal structure of the grains of
the crack initiation and propagation mechanism also has been qualitatively analyzed.

PARTICLE CRUSHING STRENGTH

Uniaxial compression tests are used to study the mechanical behavior of these particles. In
this test, the particle is compressed between two rigid flat plates. Crushing strength of the
particles depends on the failure force and the initial dimension of the particle along loading. The
method proposed by Oka and Hiramatsu is employed in this study to find out the crushing
strength of each particle (Hiramatsu and Oka 1966).

𝐹𝑓
𝜎 = 0.9 (1)
𝑑𝑜2

The strength of some manmade materials like steel is constant because they have uniform
structures that is not dominated by the presence of flaws. However, natural materials like soil
have structures dominated by flaws. Therefore, instead of a single number as the strength value,
a statistical distribution called Weibull distribution has been widely used to represent the
statistical variability of the particle strength resulting from the presence of incipient flaws within
the particles (Weibull 1939, 1951, Lobo-Guerrero and Vallejo 2006). The probability of survival
(Ps) of a brittle material with volume (Vo) under stress (σ) can be described by the two-parameter
Weibull distribution function shown in Equation (2).

𝜎 𝑚
𝑃𝑠 (𝑉𝑜 , 𝜎) = 𝑒𝑥𝑝 [− ( ) ] (2)
𝜎𝑜

Using the uniaxial compression tests, the probability of survival at crushing stress σ is
calculated in Equation (3) as:

𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑝𝑎𝑟𝑡𝑖𝑐𝑙𝑒𝑠 𝑐𝑟𝑢𝑠ℎ𝑒𝑑 𝑎𝑡 (𝑆>𝜎)


𝑃𝑠 (𝑉𝑜 , 𝜎) = (3)
𝑡𝑜𝑡𝑎𝑙 𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑡𝑒𝑠𝑡𝑒𝑑 𝑝𝑎𝑟𝑡𝑖𝑐𝑙𝑒𝑠

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 567

Taking two times the logarithm from Equation (2) results in a linear relationship between the
parameters which can be derived from the uniaxial compression tests. A least square linear
regression can be conducted to determine the Weibull parameters (σo, m), described by Equation
(4).

1
ln[ln ( )] = 𝑚𝑙𝑛𝜎 − 𝑚𝑙𝑛𝜎𝑜
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

(4)
𝑃𝑠 (𝑉𝑜 ,𝜎)

In this expression σo is the scale parameter, called characteristics stress, as an indication of


the mean crushing strength of the particles and m is the shape parameter, called Weibull
modulus, as an indication of the variability of the particle crushing strength where a higher
modulus means less variability in the crushing strength (Weibull 1951, Bolton and McDowell
1997).

MATERIAL AND EXPERIMENTAL SETUP

Materials used in this study include bryozoan particles which are one of the main constituent
particles of a batch of carbonate sands from Puerto Rico. These particles are slender tubular
fragments of bryozoan skeletal remains (particle size 1-2 mm) that are shown in Figure 1. SEM
and XRD analysis on the particles revealed that these particles are mainly made of aragonite
(>98%) which have needle-like crystals as one of the polymorphs of calcium carbonates.

Figure 1. optical image of typical bryozoan particles used in this study (a-top), SEM image
of bryozoan showing aragonite crystals (a-bottom), experimental setup for uniaxial
compression test (b-top), Micro-computed tomography scanning setup (b-bottom)

37 uniaxial compression tests have been conducted on the bryozoans by a micro tensile tester
device from Deben, UK. The device was used in compression mode and the uniaxial
compression tests were conducted under a Hirox KH-8700 optical microscope to record the
deformation of particle during compression. The experimental setup can be seen in Figure 1(b).

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 568

Displacement-controlled compression tests were conducted at a constant displacement rate of 0.1


mm/min until cracks fully developed and the particle broke and was not able to sustain any load.
Failure force for each grain is obtained from the load-displacement curve as the peak force and
the initial particle size between the loading plates is measured using Fiji/Image J software
(Schindelin et al. 2012) from the optical images.
To investigate the internal structure, particles were scanned by Zeiss Xradia 510 Versa at
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

50kV, 4W, with no filter, using 4X objective and 1601 projections with 360 degrees of rotation.
The scans have been processed in Dragonfly software 2020.1 (Object Research Systems,
Montreal, Canada).
To quantify the internal structure, image processing techniques have been applied on the
acquired micro-CT scans. Image processing techniques can be found in several studies (Kikkawa
et al. 2013, Kong and Fonseca 2018, Beemer et al. 2022). The gray-scale CT scans have been
segmented into solid and pore phase by using Otsu’s method. This method is a global intensity
thresholding method applicable to images with bimodal histograms. In this technique, the
threshold is determined by maximizing the intensity variance of inter-classes which correspond
to minimizing the intra-class intensity variance (Otsu 1979).
After, segmenting the scans, the internal porosity of the particle is measured by dividing the
volume of pore pixels to the total volume of particle pixels. The full statistical analysis of the
pore structure including several pore descriptors such as pore size, pore orientation and pore
connectivity can be done by pore network modeling which is not the focus here. More details on
the method and its applications on carbonate sand particles can be found in (Mohtashami et al.
2022).

RESULTS AND DISCUSSION

The internal structure of bryozoans has been extracted from the micro-CT scans. The scans
revealed that the internal structure of bryozoans is mainly include large pores that have been
connected through large and long channels that goes through the whole particle. Therefore, there
is a network of pores in the bryozoans than isolated pores which make their fracture mechanism
complicated. The scams have been segmented into pore and solid phase and porosity is measured
as 33%.

Figure 2. Bryozoan particle along with segmented micro-CT scans in three perpendicular
planes; (b) Z plane, (c) Y plane, (d) X plane, red color represents pore phase and green
color represents solid phase

From the load-displacement curves derived from the uniaxial compression tests, it showed
that most bryozoan particles (34 out of 37) go through a progressive fracture with only 3 grains
show sudden load drops with no more ability to sustain loads. While, silica sand grains have
been reported to show brittle fracture, bryozoan particles showed more of a ductile behavior.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 569

Also, for these particles fracture and formation of cracks can be followed easily under the
microscope. Among 34 with the progressive fracture, 13 particles gain strength upon loading.
This can be attributed to the pore collapse and deterioration that can occur inside the particle that
fills up other pores as the result of loading leaving a solid matrix of material. The load-
displacement corresponding to the mentioned behavior are shown in Figure 3.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Three types of load-displacement curves observed for bryozoan particles

Weibull analysis has been performed on the crushing strength data and Weibull modulus and
characteristics stress are derived as 2.18 and 2.95 MPa respectively. The result of fitting Weibull
distribution to the crushing strength data are shown in Figure 4. The Weibull modulus for silica
sand has been reported to be around 3 in the literature. The lower Weibull modulus of bryozoans
can be attributed to the internal structure of these particles.

Figure 4. Probability of survival of the particles measured in uniaxial compression test data
and the fitted Weibull

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 570

Bryozoans have a small characteristics strength comparing the reported values for carbonate
sands. This can be attributed to the high porosity of these particles and the interconnected pores
and pore channels which contributed to the easy crack propagation through the material. Some of
the reported values in the literature are summarized in Table 1.

Table 1. Summary of Weibull parameters for different selected silica and carbonate sands
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Characteristic
Particle Size Particle Weibull
Material Strength
(mm) Shape Modulus
(MPa)
0.25-0.3 1.82 110.87
Silica Sand (Nakata et al.
0.6-0.7 Angular 2.17 72.18
2001)
1.4-1.7 3.04 30.96
Ottawa Silica Sand (Cil Rounded
0.599-0.853 3.26 137
and Alshibli 2012)
Carbonate Sand, 3.58 1.76
Australia (Beemer et al 1
5.11 26.57
2019)
5-8.45 1.79 6.54
Coral Sand, Philippines Relatively
8.45-10.50 2.27 6.26
(Xuehui et al. 2020) rounded
10.50-14.50 2.00 5.12
1-2 1.66 23.62
Carbonate Sand, South
2-5 1.96 8.40
of China (He et al. 2022)
5-10 3.29 3.49
Carbonate Sand, Playa
Santa, Puerto Rico (This 1-2 Tubular 2.18 2.95
study)

Based on our observations on the particle after loading under optical microscope, there were
main cracks either along the loading direction or along the longer direction of the particle. There
were other branched cracks that were propagated along the pores. A few of the particles are
shown in Figure 3. Fracture surfaces of the broken pieces were investigated using a scanning
electron microscope as shown in Figure 5. (d). It was observed that the fracture surfaces cut
through the intragrain pore network following zones of weakness within the grain.
To investigate the crack initiation and propagation in bryozoan particles, micro-CT scans of a
loaded particle have been provided in Figure 5 (a-c). For this grain, the loading was stopped at
the first load drop and the grain was scanned then to investigate the generation of the crack and
its geometry. The 3D reconstruction of the loaded grain using micro-CT scans along with the
optical images captured during loading showed that in this particle, in contrast to what was
expected, the first crack was not initiated at the points of contacts between loading plates and the
particle. It was generated inside the particle presumably at the internal pores as points of stress
concentrations. Upon loading, the crack propagated along the loading tips up to a point (O) and
propagated as the second crack along the loading direction from that point Figure 5 (e-g). The
crack path followed the internal pores. While further studies are needed to conclusively
investigate the effect of the complex internal structure of carbonate sands on their breakage
behavior, our results provide evidence on contributing role of the internal structure on the
fracture behavior of the particles including the crack initiation and propagation mechanism.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 571
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 5. micro-CT scans of a bryozoan particle after loading (top row), load was applied
perpendicular to ZY plane, loading direction is shown by red arrow

CONCLUSIONS

In this study we took an experimental approach to investigate the crushing strength and
breakage behavior of biogenic carbonate sand particles from Puerto Rico. Uniaxial compression
tests and micro computed tomography are employed to analyze the breakage behavior and
internal structure of the particles. The concluding remarks of this paper can be summarized as
below:
• The internal structure of bryozoans includes considerable number of pores. Porosity for a
bryozoan particle is calculated as 33%. The pore structure of these grains includes a
network of inter-connected pores which facilitates the crack propagation inside these
particles
• The crushing strength of bryozoan particles follows Weibull distribution. The
characteristic strength for these particles is measured as 2.5 MPa which is considerably
lower than the values reported for silica sands. This value is considered small as well
comparing the similar value reported for other carbonate sand particles. Furthermore,
Weibull modulus of carbonate sand is lower indicating a higher variability of its crushing
strength. This can be attributed to the complex internal structure of carbonate sand
• This study made an effort in understanding the contributing role of the internal structure
of carbonate sands on their fracture behavior. Based on the acquired results, internal
pores can be considered as the points of stress concentration where cracks initiated.
• The results emphasize on the important role of the internal structure of carbonate sands
on their breakage behavior and the internal structure should be considered as one of the
influential factors in the models and design schemes. The data also adds to the current
database of the studies investigating the mechanical behavior of carbonate sands at grain-
scale level.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 572

REFERENCES

Airey, D. W., Randolph, M. F., and Hyden, A. M. (1988). “The Strength and Stiffness of Two
Calcareous Sands.” Proc. of the Int. Conf. on Calcareous Sediments, Perth, vol. 1, no. 2pp.
43–50.
Beemer, R. D., Bandini-Maeder, A., Shaw, J., and Cassidy, M. J. (2019a). “Volumetric Particle
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

size Distribution and Variable Granular Density Soils.” Geotechnical Testing Journal, 43(2),
517–533.
Beemer, R. D., Li, L., and Leonti, A. (2022). “Comparison of 2D Optical Imaging and 3D
Microtomography Shape Measurements of a Coastal Bioclastic Calcareous Sand.” Journal of
Imaging, 8(3), 72.
Beemer, R. D., Sadekov, A., and Lebrec, U. (2019b). “Impact of Biology on Particle Crushing in
Offshore Calcareous Sediments.” Geo-Congress 2019: Geotechnical Materials, Modeling,
and Testing, American Society of Civil Engineers Reston, VA, vol. 1341, no. 2016pp. 640–
650.
Bolton, M. D., and McDowell, G. R. (1997). “Clastic Mechanics.” IUTAM Symposium on
Mechanics of Granular and Porous Materials, Springer, Dordrecht, pp. 35–46.
Brandes, H. G. (2011). “Simple Shear Behavior of Calcareous and Quartz Sands.” Geotechnical
and Geological Engineering, 29(1), 113–126.
Cil, M. B., and Alshibli, K. A. (2012). “3D Assessment of Fracture of Sand Particles using
Discrete Element Method.” Geotechnique Letters, 2(7–9), 161–166.
Coop, M. R. (1988). “Particle Crushing of Carbonate Sands.” Proceeding of the 15th Australian
Conference on the Mechanics of Structures and Materials, Perth, vol. 2pp. 875–876.
Coop, M. R. (1990). “The Mechanics of Uncemented Carbonate Sands.” Geotechnique, 40(4),
607–626.
Druckrey, A. M., and Alshibli, K. A. (2016). “3D Finite Element Modeling of Sand Particle
Fracture Based on In situ X‐Ray Synchrotron Imaging.” International Journal for Numerical
and Analytical Methods in Geomechanics, 40(1), 105–116.
Fookes, P. G. (1988). “The Geology of Carbonate Soils and Rocks and their Engineering
Characterisation and Description.” International Conference on Calcareous Sediments, pp.
787–806.
Golightly, C. R., and Hyde, A. F. L. (1988). “Some Fundamental Properties of Carbonate Sands,
Engineering for Calcareous Sediments.” Proc. Int. Conf. on Calcareous Sediments, 1988,
Perth, vol. 1, no. 2pp. 69–78.
Hassanlourad, M., Salehzadeh, H., and Shahnazari, H. (2008). “Dilation and Particle Breakage
Effects on the Shear Strength of Calcareous Sands based on Energy Aspects.” International
Journal of Civil Engineering, 6(2), 108–119.
He, Y., Cai, G., Gao, L., and He, H. (2022). “Effect of Particle Size and Constraint Conditions
on Single Particle Strength of Carbonate Sand.” Sensors, 22(3), 1–17.
Hiramatsu, Y., and Oka, Y. (1966). “Determination of the Tensile Strength of Rock by a
Compression Test of an Irregular Test Piece.” International Journal of Rock Mechanics and
Mining Sciences and, 3(2), 89–90.
Hyodo, M., Aramaki, N., Itoh, M., and Hyde, A. F. L. (1996). “Cyclic Strength and Deformation
of Crushable Carbonate Sand.” Soil Dynamics and Earthquake Engineering, 15(5), 331–336.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 573

Kikkawa, N., Orense, R. P., and Pender, M. J. (2013). “Observations on Microstructure of


Pumice Particles using Computed Tomography.” Canadian Geotechnical Journal, 50(11),
1109–1117.
Kong, D., and Fonseca, J. (2018). “Quantification of the Morphology of Shelly Carbonate Sands
using 3D Images.” Geotechnique, 68(3), 249–261.
Kuang, D., Long, Z., and Guo, R. (2021). “Experimental and Numerical Study on the
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Fragmentation Mechanism of a Single Calcareous Sand Particle under Normal


Compression.” Bulletin of Engineering Geology and the Environment, 1–14.
Kwag, J. M., Ochiai, H., and Yasufuku, N. (1999). “Yielding Stress Characteristics of Carbonate
Sand in Relation to Individual Particle Fragmentation Strength.” Engineering for Calcareous
Sediments, 1, 79–86.
Li, H. Y., Chai, H. W., and Xiao, X. H. (2020). “Fractal Breakage of Porous Carbonate Sand
Particles: Microstructures and Mechanisms.” Powder Technology, 363, 112–121.
Lobo-Guerrero, S., and Vallejo, L. E. (2006). “Application of Weibull statistics to the tensile
strength of rock aggregates.” Journal of geotechnical and geoenvironmental engineering,
132(6), 786–790.
Lv, Y., Li, X., and Wang, Y. (2020). “Particle Breakage of Calcareous Sand at High Strain
Rates.” Powder Technology, 366, 776–787.
Ma, L., Li, Z., and Wang, M. (2019). “Effects of Size and Loading Rate on the Mechanical
Properties of Single Coral Particles.” Powder Technology, 342, 961–971.
Al Mahbub, A., and Haque, A. (2016). “X-ray Computed Tomography Imaging of the
Microstructure of Sand Particles Subjected to High Pressure One-dimensional Compression.”
Materials, 9(11), 7–15.
McDowell, G. R., and Harireche, O. (2002). “Discrete Element Modelling of Yielding and
Normal Compression of Sand.” Geotechnique, 52(4), 299–304.
Mohtashami, E., Olgun, C. G., Wu, C., and Selly, T. (2022). “Intragrain Pore Structure of
Carbonate Sands.” Geo-Congress 2022, pp. 194–203.
Morales-velez, A. C., Baxter, C. D. P., Pando, M. A., and Anderson, J. B. (2015). “Comparison
of the Cyclic Resistance of a Calcareous Sand Deposit from Puerto Rico from Seismic
Dilatometer (SDMT) and Seismic Cone Penetration Tests (SCPTu).” 3rd International
Conference on the Flat Dilatometer, 7.
Nakata, Y., Hyodo, M., and Hyde, A. F. L. (2001). “Microscopic Particle Crushing of Sand
Subjected to High Pressure One-dimensional Compression.” Soils and Foundations, 41(1),
69–82.
Orense, R. P., Pender, M. J., Hyodo, M., and Nakata, Y. (2013). “Micro-mechanical Properties
of Crushable Pumice Sands.” Géotechnique Letters, 3(2), 67–71.
Otsu, N. (1979). “Threshold Selection Method From Gray-Level Histograms.” IEEE Trans Syst
Man Cybern, SMC-9(1), 62–66.
Porcino, D., Caridi, G., and Ghionna, V. N. (2008). “Undrained Monotonie and Cyclic Simple
Shear Behaviour of Carbonate Sand.” Geotechnique, 58(8), 635–644.
Poulos, H. G., Uesugi, M., and Young, G. S. (1982). “Strength and Deformation Properties of
Bass Strait Carbonate Sands.” Geotechnical Engineering, 13(2), 189–211.
Rasouli, M. R., Moradi, M., and Ghalandarzadeh, A. (2021). “Effects of Initial Static Shear
Stress Orientation on Cyclic Behavior of Calcareous Sand.” Marine Georesources &
Geotechnology, 39(5), 554–568.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 574

Saeidaskari, J., Alibolandi, M., and Azizkandi, A. S. (2020). “Undrained Monotonic and Cyclic
Behavior of Qeshm Calcareous Sand.” Marine Georesources & Geotechnology, 1–14.
Salem, M., Elmamlouk, H., and Agaiby, S. (2013). “Static and Cyclic Behavior of North Coast
Calcareous Sand in Egypt.” Soil Dynamics and Earthquake Engineering, 55, 83–91.
Sandoval, E. A., and Pando, M. A. (2012). “Experimental Assessment of the Liquefaction
Resistance of Calcareous Biogenous Sands.” Earth Sciences Research Journal, 16(1), 55–63.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Schindelin, J., Arganda-Carreras, I., and Frise, E. (2012). “Fiji: An Open-source Platform for
Biological-image Analysis.” Nature methods, 9(7), 676–682.
Weibull, W. (1939). “A Statistical Theory of Strength of Materials.” IVB-Handl., 151, 1–45.
Weibull, W. (1951). “A Statistical Distribution Function of Wide Applicability.” Journal of
Applied Mechanics, 18(3), 293–297.
Xiao, Y., Liu, H., and Chen, Q. (2017). “Particle Breakage and Deformation of Carbonate Sands
with Wide Range of Densities During Compression Loading Process.” Acta Geotechnica,
12(5), 1177–1184.
Xuehui, W., Yuanqiang, C., and Sifa, X. (2020). “Effects of Size and Shape on the Crushing
Strength of Coral Sand Particles under Diametral Compression Test.” Bulletin of Engineering
Geology and the Environment, 1–11.
Zhang, J. M., Duan, M. D., Wang, D. L., and Zhang, Y. (2019). “Particle Strength of Calcareous
Sand in Nansha Islands, South China Sea.” Advances in Civil Engineering Materials, 8(1),
355–364.
Zhao, B., Wang, J., and Andò, E. (2020). “Investigation of Particle Breakage under One-
dimensional Compression of Sand using X-ray Microtomography.” Canadian Geotechnical
Journal, 57(5), 754–762.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 575

An Evaluation of Incremental, Constant Rate of Strain and Constant Pressure Ratio


Consolidation Testing

Ryan Lavorati, P.E.1; William Marr, Ph.D., P.E., NAE2; Kaveh Zehtab3; Salim Werden4;
and John Christian, Ph.D., P.E., NAE5
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

1
Assistant Project Manager, Geocomp. Email: rlavorati@geocomp.com
2
CEO, Geocomp. Email: wam@geocomp.com
3
Testing and Automation Leader, Geocomp. Email: kzehtab@geocomp.com
4
Testing and Automation Developer, Geocomp
5
Professor, Univ. of Massachusetts Lowell

ABSTRACT

Consolidation testing is an essential part of geotechnical testing, particularly for settlement


analyses and staged construction. Two of the most common tests are incremental loading and
constant rate of strain tests. Incremental loading tests hold a particular load over a 24-h duration,
typically taking about a week to run, providing limited stress-strain data points, and measures
secondary compression characteristics. Constant rate of strain tests holds the rate of strain
constant throughout the test, can take about two to three days to run depending on the strain rate,
provides a near continuous stress-strain diagram, and limited measurements of secondary
compression. This paper presents an alternative to these consolidation testing methods. First
proposed in the early 1980s, the constant pressure ratio consolidation test varies the strain rate to
maintain the ratio of excess pore pressure to total stress constant throughout the test. To maintain
this ratio, a computer is required to loop the strain rate to hold the ratio of the excess pore
pressure to total stress to the desired level. An experimental program consisting of 16 re-
sedimented Boston blue clay compared the results of the three consolidation test methods. The
experimental program also examined conditions unique to the constant pressure ratio method,
namely the update time interval. The results for the constant pressure ratio test indicate the test
finishes within hours, often taking more time to set up and saturate the sample up than to run the
test. The constant pressure ratio test provides similar a near continuous stress-strain and similar
coefficients of consolidation across the three tests. The constant pressure ratio test provides
limited measurements of secondary compression, like the constant rate of strain test. The update
time interval tells the software how often to loop and update the strain rate to maintain the
pressure ratio. This paper demonstrates that the constant pressure ratio consolidation test
provides results faster than the two common consolidation tests, while providing similar results
to both tests.

INTRODUCTION

The purpose of this paper is to compare the results of the constant pore pressure ratio
consolidation method to the constant rate of strain and incremental loading methods. The
constant pore pressure ratio method is the most recent method developed, while the constant rate
of strain and incremental loading methods both have been around for several decades with
ASTM standards.
The equipment used includes the commercial testing machines adapted independently for
testing incremental load consolidation and constant rate of strain consolidation. The equipment

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 576

setup for the constant pressure ratio test used a similar setup as the constant rate of strain
consolidation. The notable differences included the use of a second pore pressure transducer line
attaching to the cell chamber for a direct measurement of excess pressure measurements. The
material used in this lab testing program was resedimented Boston Blue Clay. This material is a
moderate plasticity clay with well-known properties (House, 2009).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

BACKGROUND

The Incremental Loading test is the original testing method for determining consolidation
characteristics of clay soils (Terzaghi, 1923). These parameters are developed using stress-strain
and strain-time plots to determine the parameters graphically.
ASTM D2435 has two methods for determining how long the load increment should be left
on the sample (American Society of Testing and Materials, 2011). Method A prescribes that the
load increment should be left on the sample for 24 hours. Method B prescribes that the load
increment should be left on the sample until the point of end of primary consolidation with a
predetermined amount of time after that point. Depending on the number of load increments,
either method will result in the test taking days to perform.
The constant rate of strain test was first pioneered by Smith and Wahls in 1969 (Smith &
Wahls, 1969). The analysis method from the Smith and Wahls paper used a solution that applies
to the steady state condition for linear material behavior. The constant rate of strain test was
improved in 1971 at the Massachusetts Institute of Technology (Wissa, Christian, Davis, &
Heiberg, 1971) to include the solution for steady state and transient conditions, and for linear and
non-linear material behavior. Following the Wissa paper in 1971, the constant rate of strain test
started to become a generally accepted consolidation testing method. There are several
commercially available constant rate of strain test machines, and some custom designs are
possible. ASTM D4186 provides the minimum requirements for performing the constant rate of
strain test. Although the test has been described as a constant rate of strain test since its
conception, the test is really a constant rate of deformation.
The development of the constant pore pressure ratio test started as early as the 1980s in
Norway. The first published paper involving the constant pore pressure ratio test was in 1981
(Janbu, Tokheim, & Senneset, 1981). The first published paper discusses the consolidation
theory behind the constant pore pressure ratio test, some limitations, and a limited testing
program.
The Janbu et. al, 1981 paper starts with the data acquisition, interpretation, and plotting of the
results that are completely automated. That paper explains that the automation is required due to
the need to monitor the changes in pore pressures actively. The changes in pore pressure are
needed so the machine can respond to maintain the target pore pressure ratio. The known
parameters recorded were the applied vertical load, sample compression, and the pore pressure at
the base of the sample, as well as the rate with respect to time for developing these parameters.
Based on these recorded parameters and the consolidation theory developed in the Janbu et. al,
1981, the several equations are presented to evaluate the constant pressure ratio test from a
theoretical perspective. The derived equations are listed below.

𝑢̇ 𝑏
𝜆= (1)
𝑞̇

(1 − 𝜆) ∗ cosh 𝑎 = 1 (2)

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 577

𝑞̇ 𝐻
𝑀 = 𝛼𝑀 (3)
𝛿̇

tanh 𝑎
𝛼𝑀 = (4)
𝑎
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

𝛾𝑤 𝐻𝛿̇
𝑘 = 𝛼𝑘 (5)
2𝑢𝑏

2(cosh 𝑎−1)
𝛼𝑘 = (6)
𝑎∗sinh 𝑎

𝑞̇ 𝐻 2
𝑐𝑣 = 𝛼𝑐 (7)
2𝑢𝑏

2(cosh 𝑎−1)
𝛼𝑐 = (8)
𝑎2 ∗cosh 𝑎

where,
λ = pore pressure ratio
M = constrained modulus
H = sample height
γw = unit weight of water
𝑞̇ = applied load rate
𝛿̇ = displacement rate
a = coefficient
q = applied load
δ = displacement
ub = pore pressure at the base
𝑢̇ b = pore pressure rate at the base
Equations 1 through 8 can be found in the Janbu et. al, 1981 paper. Equation 1 provides the
magnitude of the pore pressure ratio during the test as the ratio between the change in pore
pressure rate at the base of the consolidation test to the change in total stress. Equation 2 is an
equation needed to develop the coefficient “a” for Equations 4, 6, and 8. Equations 3, 5, and 7
provide the calculation for constrained modulus, permeability, and coefficient of consolidation,
based on the previous equations as well as parameters measured during the test, such as the base
pore pressure or the displacement rate.
The Janbu et. al, 1981 paper then transitions into some tests that were performed using the
constant pressure ratio test. Some of the major findings from its tests on the constant pressure
ratio method include consistent results from constant pressure ratios varying from 0.2 to 0.5, the
magnitude of the pre-consolidation pressure appears to be the same for both the constant pressure
ratio test and constant strain rate tests if the pressure ratio is less than 35%, and the coefficient of
compressibility versus effective tress varies in a similar manner to the tangent modulus versus
effective stress.
Following the 1981 paper, either no immediate work was continued on the constant pore
pressure ratio method, or no work was published to advance this work.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 578

MATERIALS

The materials used in this assessment consisted of resedimented Boston Blue Clay. A bulk
Boston Blue Clay sample was collected from a construction site at 200 Pier Four Boulevard in
Boston, Massachusetts. The approximate depth of the samples collected was between 30 and 40
feet below the ground surface. The material was transported to the geotechnical lab and thinly
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

laid out on tarps, and allowed to air dry with fans circulating the air. Larger chunks were
manually broken down by hand or with a weighted plate. During this time, any particles
suspected of being larger than the #10 sieve were removed. Once the sample appeared to be
completely dry, the Boston Blue Clay was then ground down using a Retsch BB51 pulverizer to
100% passing the #100 standard US sieve. Material that did not pass the #100 sieve was put back
through the pulverizer. Processed Boston Blue Clay was put into 5-gallon buckets until
resedimentation.

Table 1. Index Results on Resedimented Boston Blue Clay

Parameter Range (Average)


Liquid Limit (%) 26 – 35 (31)
Plastic Limit (%) 13 – 15 (14)
Plastic Index (%) 12 – 22 (17)
Water Content (%) 25.2 – 28.8 (26.7)
Specific Gravity 2.69 – 2.72 (2.71)

EQUIPMENT AND PROCEDURES

The resedimentation process creates soil sample that can be cut and trimmed into multiple
specimens for consolidation testing. The resedimentation process involves preparing a uniform,
homogeneous workable slurry, consolidating the slurry one dimensionally in a large, aluminum
consolidometer.
Resedimentation of Boston Blue Clay was first undertaken at the Massachusetts Institute of
Technology by Bailey in 1961 (Abdulhadi, 2003). In 1982, the process of resedimentation was
improved to produce fully saturated and uniform samples with salt concentrations of 16 g/L
(Germaine, 1982).
The clay powder is gradually blended into the salt water using a hand mixer. Once a smooth,
uniform slurry is created, the slurry is poured into a large flask for de-airing. The de-airing is
performed under pressure for approximately 1 hour to remove air from the slurry. The slurry is
shaken periodically while in the de-airing flask. After the approximate hour, the slurry is
transferred to the batching chamber.
The batch chamber is made of aluminum. On one side of the batch chamber, nozzles with
connectors are found for transmitting water. The top cap of the batch chamber is also made of
aluminum, with a gasket around the edge to provide an airtight seal within the chamber and a
nozzle on top of the cap. The nozzle allows for water to drain from the top of the sample. Plastic
piping is connected to the top cap and the bottom of the batch chamber to allow double drainage.
The plastic piping extends to two containers partially filled with 16 g/L salt water. Prior to
batching, the chamber and the chamber cap are washed clean with water, allowed to dry, and
then scrubbed with WD40 oil. The base of the consolidometer is covered by a porous plastic
stone, a filter fabric, and filter paper.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 579

After the filter paper on the bottom of the chamber has been placed, the soil slurry is gently
poured into the chamber. The batching, de-airing, and pouring process is repeated until the
desired initial height of sample is achieved in the consolidometer. After the desired initial height
of sample is achieved, filter paper, filter fabric, and a porous plastic stone are gently placed on
top of the sample, with the filter paper in contact with the soil slurry. The top cap is then put into
position as close as possible to the top of the top porous stone.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

After the consolidometer set up is complete, the batch is resedimented like an incremental
consolidation test. Stress, strain, and time are recorded through the software program and saved
in a computer file. The sample is loaded incrementally up to 2000 psf and allowed to creep for
one log cycle of time past the end-of-primary consolidation time at the 2000 psf step. The sample
is then unloaded to 500 psf and allowed to relax for one log cycles of time past the end-of-
primary consolidation time at the 500 psf step. This creates an over consolidation ratio of
approximately 4. At an over consolidation ratio of 4, the lateral stress ratio is approximately 1
(Ladd R., 1965), providing isotropic effective stress conditions. A product of the sample being at
this stress state is that the shear strains during sample extrusion and trimming should be minimal
(Adams, 2008).
After the resedimented samples have been created, each sample underwent a number of tests.
All index test results were performed following their respective ASTM standards. The index
results are summarized in Table 1.
Consolidation testing generally followed their respective ASTM standards. The Incremental
Loading tests were performed following test method B of the ASTM standard (American Society
of Testing and Materials, 2011). A load increment ratio of 1 was used for each step. This load
increment ratio is typical of most testing. For this paper, the Constant Rate of Strain tests were
performed following the current ASTM standard (American Society of Testing and Materials,
2012). During the test, the sample was fixed with an upper bound and lower bound of 3% to 15%
for the excess pore pressure ratio. If the sample experienced an excess pore pressure ratio lower
than 3%, the strain rate would increase to maintain the pressure ratio of 3%. If the sample
experienced an excess pore pressure ratio higher than 15%, the strain rate would decrease to
maintain the pressure ratio at 15%. The first test was performed at a strain rate of 1%/hr and the
second test was performed at a strain rate of 2%/hr. These strain rates are consistent with what is
recommended in practice (Ladd & DeGroot, 2003) and recommended from research (Gonzalez,
2000).
The Constant Pressure Ratio Test is set up similar in set up to the Constant Rate of Strain
Test. Pore pressure ratios were selected based on literature review and experimental trial and
error. In setting up the testing procedure, six input parameters are needed. The first is the final
normal stress. This stress is the effective stress applied to the sample before it switches to the
next step. The second is the pressure ratio. This is the target pore pressure ratio during the test.
The third is the initial strain rate. This strain rate is generally based on literature and experience.
The strain rate during the test will change as the test progresses. The next components are the
equilibrium pressure ratio and the maintain time. When the step reaches the final normal stress,
the sample will target the equilibrium pressure ratio and maintain this ratio for the maintain time.
This part of the step is to smooth out the transition between steps. The last input is the update
interval. This input tells the computer how often to update the strain rate in response to the
generated excess pore pressures compared to the applied vertical stress. In this experimental
program, the inputs pressure ratio and update interval were varied experimentally.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 580

DATA ANALYSIS AND METHODS

The analysis, data reduction, and calculations used for the incremental loading, constant rate
of strain, and constant pressure ratio tests were analyzed using software developed by the
GeoComp Products division (Geocomp Products, 2018).
For the incremental loading test, the coefficient of consolidation calculated in this testing
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

program used Taylor’s method using the square root of time (Taylor, 1948). Taylor’s method
was used to calculate the coefficient of consolidation at T90 and T100.
For the constant pressure ratio test, the software was updated to provide a loop for
determining and holding the target pore pressure ratio. The software applies a load to the sample,
determines the excess pressure, calculates the excess pressure ratio, and determines whether to
increase or decrease the strain rate to reach and maintain the target excess pressure ratio.
Several consolidation parameters were not calculated by the software packages and required
user interpretation. These parameters include the pre-consolidation stress, recompression ratio,
compression ratio, and swell ratio. The pre-consolidation stress was determined using Becker’s
Strain Energy approach (Becker, Crooks, Been, & Jefferies, 1987).

RESULTS

Twelve Constant Pressure Ratio tests were performed on samples of resedimented Boston
Blue Clay. Of these twelve tests, two tests were performed at a pressure ratio of 15%, three tests
were performed at a pressure ratio of 9%, two tests were performed at a pressure ratio of 3%. For
these tests, the target pressure ratio was applied to both the loading and unloading conditions.
During these tests, the selected update interval was 55 seconds. These target pressure ratios are
based on the limits described in ASTM D4186, as well as one data set in between these limits.
After these tests were performed, three additional tests were performed varying the update
interval. The update interval for each of these tests were 30 seconds, 55 seconds, and 75 seconds.
The sample was loaded using a target pressure ratio of 15% and unloaded using a target pressure
ratio of 3%. After the update interval tests, two constant pressure ratio tests were performed at
low pressure ratio. The first test had a target pressure ratio of 0.3% from the start of the test. The
results from this test are not included in this report. This is because the system could not monitor
the pore pressures at low confining stresses. The second test was consolidated in stages at
different target pressure ratios. The sample consolidated to 4000 psf, well above the apparent
pre-consolidation stress, at a target ratio of 15%. After 4000 psf, the target pressure ratio was
0.2% for the duration of the test.
Components that were unique to constant pressure ratio testing include saturating and
inundating, de-airing the system, testing duration, and equipment requirements. The specimen
was back pressured to achieve saturation. This component impacts sample behavior during
consolidation and testing. The constant pressure ratio test requires several hours to back pressure
the test to achieve saturation, and will push the air into solution. De-airing of the system
generally consisted of removing air pockets from the top and bottom of the consolidometer and
from the pipes connected to the consolidometer. For the constant pressure ratio tests, the de-
airing of the system was required for three pipes and the top and bottom of the consolidometer.
The top and bottom of the consolidometer are de-aired in a similar manner as the constant rate of
strain set up. For the pipes of the constant pressure ratio test, one pipe is for the pump, the second
is for the low end of the pressure transducer, and the third is for the high end of the pressure

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 581

transducer. The pipes for the low- and high-end transducers are de-aired by pumping water through
the pipes to the transducer until all of the air appears to be removed from the piping system.
Once each test is ready to be tested, the duration between each test can vary significantly.
The testing duration for the incremental loading tests took approximately 4,200 minutes. For the
constant pressure ratio tests, the testing duration varied from approximately 400 minutes
(particularly at 15%) to 1,700 minutes (particularly for 3%). When comparing the incremental
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

loading tests to the constant pressure ratio tests between 3% and 15%, the constant pressure ratio
tests were approximately 2.5 to 11 times faster.

Figure 1. Stress-Strain Curve Comparison

Given that the materials are re-sedimented of the same materials, the results for the
recompression ratio, compression ratio, swell ratio, and pre-consolidation stress should be
reasonably close to each other. As shown in Table 2, the results between the parameters between
these three tests fall within a small window between each test method. The similarity between
the stress-strain relationship between the constant pressure ratio tests at target pressure ratios of
3%, 9%, and 15%, the constant rate of strain consolidation at target strain rates of 1%/hr and
2%/hr, and the incremental loading test can be seen on Figure 1.
The coefficient of consolidation was also calculated for each of the three tests. As shown in
Figure 2, the coefficient of consolidation is plotted against effective stress. The data shown on
the plot appears to have some noise. This is based on how frequently the software collected the
data from the test. It is apparent from the graph that although there is noise in the results, a clear
trend can be found. When compared against the values calculated from the constant rate of strain
test, the coefficient of consolidation values reasonably line up with the values from the constant
pressure ratio. Although the coefficient of compressibility is increasing log linearly, the increase
is very slight over a large stress range. The data likely agree with each other because the data are
being calculated almost continuously for each test method. The results for the incremental

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 582

loading tests plot approximately half a log cycle lower than the results for the constant pressure
ratio test from the time the test started to the end of the last loading step. The reason for the
discrepancy is that the constant pressure ratio test collected data nearly continuously and
calculated the results directly. The coefficient of compressibility from the incremental loading
tests were graphically determined using the square root of time method over a few steps.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Table 2. Comparison of Consolidation Results

Constant Pressure Incremental Constant Rate


Averaged Parameter
Ratio Tests Load Tests of Strain Tests
2,400 psf 2,500 psf 2,600 psf
Pre-Consolidation Stress
(114.9 kPa) (119.7 kPa) (124.5 kPa)
Recompression Ratio 0.024 0.021 0.018
Compression Ratio 0.105 0.102 0.106
Swell Ratio 0.010 0.009 0.010

Figure 2. Coefficient of Consolidation Comparison

The next comparison was with the strain rates during each test. As shown in Figure 3, the
strain rate for selected tests of constant rate of strain and constant pressure ratio is shown over
time. The stepping increase and decrease for the constant rate of strain tests is due to the
computer trying to maintain the bounds of the pressure ratio past those times. This clearly
indicates that the strain rate was not held constant over the entire test. For the constant pressure
ratio tests, a change in the logarithm of strain rate is apparent over the duration of the test. This
change in the strain rate occurs at around the same time the sample is transitioning from the
over-consolidated state to the normally consolidated state. In the lower pressure ratio tests, the
logarithm of the strain rate appears nearly constant in the normally consolidated portion of the
test.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 583
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 3. Axial Strain Rate Comparison

In addition to the comparison between testing methods, the set up for the constant pressure
ratio test requires an update time interval. The update interval is required to update the strain rate
within the data acquisition system to maintain the target pressure ratio. Ideally, the update
interval should reflect how the pore pressure of the soil sample reacts to a given strain rate. For
example, at too fast an update interval, the strain rate would become too chaotic to develop a
trend. At too slow an update interval, the sample will tend to behave more like the theory within
the constant rate of strain. Almost all the tests were conducted at an update interval of 55
seconds. Two other tests were performed at different update intervals; one at 30 seconds and one
at 75 seconds. The results of these three update intervals are shown on Figure 4. This figure
indicates that the update interval between 30 seconds and 75 seconds reasonably achieves and
maintains the same data for low plasticity clay samples.

Figure 4. Update Interval Comparison

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 584

Table 3. Comparison of Consolidation Testing Methods

Constant Pore
Testing Incremental Loading Constant Rate of Strain
Pressure Ratio Test
Component Test (IL) Test (CRS)
(CPR)
Lowest Moderate upfront cost High upfront cost
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

- Load frame and - Load frame, - Load frame,


displacement sensor displacement sensor, displacement sensor,
Initial
- Computer optional fluid pumps, piping, and fluid pumps, piping,
Investment
pressure transducer and two pressure
- Computer optional transducers
- Computer required
Simplest Moderate complexity High complexity
- Load and displacement - One pressure transducer - Two pressure
sensors/calibration only and IL transducers and IL
Preparation
- Inundation only sensors/calibration sensors/calibration
of
- Without computer, - De-airing needed - De-airing needed
Equipment
technician needs to - Without computer, - Familiarity with
monitor and record data technician needs to computer software
monitor and record data and hardware
Days to weeks Days Hours
- Low turnaround time for - Moderate turnaround - Fast turnaround time
results time for results for results
- Technician needed to add - Time is needed for - Time is needed for
water to test periodically saturating sample saturating sample
Testing
- Higher potential for - Without computer, - Computer monitors
Duration
“accidental” disturbance technician at machine and collects data
- Without computer, periodically - Strain rate varies to
technician always at - Strain rate unknown to maintain target ub/σv
machine achieve ub/σv ≈ 3% to
15%
Limited data points Nearly continuous data Nearly continuous data
- cv determined graphically - cv determined - cv determined
- Displacements at T100 mathematically mathematically
determined graphically - Effective stresses - Effective stresses
Post-Test - mv based on T100 and no calculated directly calculated directly
Results creep - Pre-consolidation stress - Pre-consolidation
σ-ε-cv - Pre-consolidation stress and log-stress/strain stress and log-
and log-stress/strain ratios determined from a stress/strain ratios
ratios determined from nearly continuous curve determined from a
limited data points nearly continuous
curve
Post-Test - Provides direct method to - Provides indirect method - Provides indirect
Results determine cαε to estimate cαε only if method to estimate cαε
ε-time tested only if tested

CONCLUSIONS

The incremental loading test has a large history of research compiled since its inception in
the 1920s. The constant rate of strain test also has been developing a large research database

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 585

since its inception, especially for its faster turnaround time and nearly continuous results. The
constant pore pressure ratio test was developed in the early 1980s and offered arguably the
fastest turnaround time with accurate and repeatable results. However, because the constant pore
pressure ratio test required a computer to run the test, it was not generally accepted and was not
followed up. The constant pressure ratio test method would not resurface for a few decades later
to be mentioned in passing. This paper shows and supports the results mentioned in the 1981
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Janbu.
A summary comparing the three different consolidation methods is described in Table 3. This
summary compares the initial investment, preparation of equipment, testing duration, post test
results for stress, strain, and coefficient of consolidation, and the post test results for strain and
time. Generally, comparing the incremental loading test to constant rate of strain test to constant
pore pressure ratio test, the initial investment and preparation of equipment increase with testing
complexity. The testing duration decreases, with the constant pore pressure ratio test being the
shortest duration test and the incremental loading test being the longest test. The drawback on
testing duration is that incremental loading tests provide multiple, direct measurements of
secondary compression parameters and the constant pore pressure ratio test would not.
Parameters developed from consolidation tests, including stress, strain, and coefficient of
consolidation, are generally similar between the three test methods.

REFERENCES

Abdulhadi, N. O. (2003). An Experimental Investigation into the Stress-Dependent Mechanical


Behavior of Cohesive Soil with Application to Wellbore Instability. Civil Engineering.
Cambridge: Massachusetts Institute of Technology, Thesis.
Adams, A. L. (2008). Laboratory Evaluation of the Constant Rate of Strain and Constant Head
Techniques for Measurement of the Hydraulic Conductivity of Fine Grained Soils.
Cambridge: Massachusetts Institute of Technology, Thesis.
ASTM. (2011). ASTM D2435-11 - Standard Test Methods for One-Dimensional Consolidation
Properties of Soils Using Incremental Loading. West Conshohocken: American Society of
Testing and Materials.
ASTM. (2012). ASTM D4186-12e1 - Standard Test Method for One-Dimensional Consolidation
Properties of Saturated Cohesive Soils Using Controlled-Strain Loading. West
Conshohocken: ASTM International.
Becker, D. E., Crooks, J., Been, K., and Jefferies, M. (1987). Work as a Criterion for
Determining in situ and yield stresses in clay. Canadian Geotechnical Journal, 24(4), 549-
564.
Geocomp Products. (2018). Geocomp Lab Products Catalogue. Retrieved from Geocomp
Products: http://www.geocomp.com/files/Prod_Specs/english/Geocomp-Lab-Products-
Catalogue.pdf.
Germaine, J. T. (1982). Development of the Directional Shear Cell for Measuring Cross
Anisotropic Clay Properties. Cambridge: Massachusetts Institute of Technology, Thesis.
Gonzalez, J. (2000). Experimental and Theoretical Investigation of Constant Rate of Strain
Consolidation. Cambridge: Massachusetts Institute of Technology, Thesis.
House, R. D. (2009). A Comparison of the Behavior of Intact and Resedimented Boston Blue
Clay (BBC). Cambridge: Massachusetts Institute of Technology, Thesis.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 586

Janbu, N., Tokheim, O., and Senneset, K. (1981). Consolidation Test with Continuous Loading.
Proceedings of the 10th International Conference on Soil Mechanics and Foundation
Engineering, 1, 645-654.
Ladd, C. C., and DeGroot, D. J. (2003). Recommended Practice for Soft Ground Site
Characterization: Arthur Casagrande Lecture. 1, pp. 1-55. Cambridge: 12th Panamerican
Conference on Soil Mechanics and Geotechnical Engineering.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Ladd, R. (1965). Use of Electrical Pressure Transducers to Measure Soil Pressure.


Massachusetts Institute of Technology, Soil Mechanics Division. Cambridge: Defense
Technical Information Center. Retrieved from
https://apps.dtic.mil/docs/citations/AD0636446.
Smith, R. E., and Wahls, H. E. (1969, March). Consolidation under Constant Rate of Strain.
Journal of the Soil Mechanics and Foundations Division, 95(SM2), 519-539.
Taylor, D. (1948). Fundamentals of Soil Mechanics. New York: John Wiley & Sons, Inc.
Terzaghi, K. (1923). Die Berechnung der Durchlassigkeitsziffer des Tones aus dem Verlauf der
Hydrodynamischen Spannungsercheinungen. 132(3/4), 125-138. Akademie der
Wissenschaften, Vienna.
Wissa, A., Christian, J. T., Davis, E. H., and Heiberg, S. (1971, October). Consolidation Testing
at Constant Rate of Strain. Journal of Soil Mechanics and Foundations Division, 97(10),
1393-1413.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 587

Numerical Analyses of a Landslide in the Sensitive Saint Adelphe Clay

Tyler J. Oathes, Ph.D., M.ASCE1; and Ross W. Boulanger, Ph.D., F.ASCE2


1
Dept. of Civil and Environmental Engineering, Rutgers Univ. Email: tyler.oathes@rutgers.edu
2
Dept. of Civil and Environmental Engineering, Univ. of California, Davis.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Email: rwboulanger@ucdavis.edu

ABSTRACT

This paper presents nonlinear dynamic analyses (NDAs) of the 1988 Saint Adelphe landslide
using FLAC 8.1 with the user defined constitutive models PM4Silt and PM4Sand. The analyses
utilized previously characterized and developed laboratory and field data, ground motions, and
site conditions. A method was introduced for modeling large sensitivities and high peak strengths
using the PM4Silt constitutive model. Calibration of the constitutive model and NDAs of the
landslide are presented. The results of the analyses indicate that the numerical methods used
could reasonably approximate the observed failure. The results also highlighted the need for the
development of models which can directly account for high peak undrained strength ratios and
large strength loss in analyses. The implications of these results on practice are discussed.

INTRODUCTION

A landslide in the Saint Adelphe municipality (about 85 km west of Quebec City) was
triggered by the magnitude (Mw) 5.9 1988 Saguenay earthquake in Canada (Abdellaziz et al.
2020). The landslide was approximately 65 m wide and 45 m long and displaced approximately
15,000 m3 of soil. Fig. 1 presents provides an aerial (Fig. 1a) and ground view (Fig. 1b) of the
landslide after the earthquake. The landslide mobilized primarily through a highly sensitive, soft
clay layer. Three in-situ geotechnical investigations were performed after the failure, one
immediately after the failure in 1988, the second in 1993, and the final in 2015. Abdellaziz et al.
(2020) summarized the in-situ investigations, presented site investigation and laboratory testing
data, and performed numerical analyses of the landslide. The simulations herein utilized the lab
and field data, the interpreted cross section, and ground motions from Addellaziz et al. (2020).
Additional details from the prior studies is presented in the relevant sections below.
The modeling of post-peak strength loss in sensitive clays with high peak strength ratios is
complicated by limitations in constitutive models and numerical solution schemes. Post-peak
strength loss in numerical analyses yields a mesh dependent solution as strains concentrate in
shear bands that are controlled by the size of the mesh. A decreasing mesh size increases the
observed strength loss which can alter the predicted deformations of a system. This is often
accounted for by incorporating a length scale wherein strength loss is scaled according to the
accumulated shear strain (e.g., Kiernan and Montgomery 2018). The introduction of
viscoplasticity can provide a measure of regularization (e.g., Oathes 2022), although limitations
in numerical methods often do not allow the coupling of viscous and dynamic behaviors (e.g.,
FLAC incompatibility of creep and dynamic analyses). Further, some advanced constitutive
models are unable to directly capture both high peak strength ratios and large sensitivities. This
can occur in critical state based models if the high peak strength ratios are mobilized at nearly
critical state. There remains a need to develop methods that directly address these modeling
limitations and better approximate the underlying mechanisms of strength loss. This paper uses a

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 588

pragmatic methodology to indirectly account for this behavior using a critical state compatible,
bounding surface plasticity model used in practice.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 1: Saint Adelphe landslide after 1988 Sagenuay earthquake: (a) aerial view and (b)
ground view (Abdellaziz et al. 2020)

This paper presents nonlinear dynamic analyses (NDAs) of the 1988 Saint Adelphe landslide
using FLAC 8.1 with the user defined constitutive models PM4Silt and PM4Sand. The analyses
utilized laboratory and field data, ground motions, and site conditions from Abdellaziz et al.
(2020). A method was introduced for modeling large sensitivities and high peak strengths using
the PM4Silt constitutive model. The calibration of PM4Silt is presented along with NDAs of the
landslide. The results of the analyses indicated that the numerical methods used could reasonably
approximate the observed failure. The results indicate this method provides a short-term indirect
way to account for high peak strengths and large sensitivities using the PM4Silt constitutive
model. The results also highlighted the need for the development of models which can directly
account for high peak strength ratios and large strength loss in analyses. The implications of
these results on practice are discussed.

Stratigraphy, Material Characterization, and Calibration

The slope was approximately 26 degrees at the toe and decreased to approximately 6 degrees
at the head of the slope. The slope contained an overconsolidated clay crust underlain by a soft,
sensitive clay layer which is underlain by a dense till layer. The crust was approximately 1-2 m
thick, and the soft clay ranged from approximately 5-9 m thick. The ground water table was at a
depth of approximately 1 m. Fig. 2 shows the stratigraphy and ground water surface as
interpreted by Abdellaziz et al. (2020).

Figure 2: Stratigraphy and water table of the slope prior to the earthquake as interpreted
by Abedellaziz et al. (2020) and as implemented in FLAC analyses

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 589

Clay Characterization and Calibration

The engineering properties of the soft clay layer were presented in Abdellaziz et al. (2020).
The soft clay had a plasticity index (PI) between 10 and 27, a liquidity index (LI) between 1 and
4, and a fines content (FC) of 70%. The sensitivity (St) was found to be between 20 and 533 from
field vane and fall cone tests. The OCR was 3 at a depth of 3m and decreased to 2 at a depth of
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

10 m. The soft clay layer was modeled using PM4Silt. The parameters were: su,cs/′vc = 0.5, hpo =
100, ho = 0.85, cz = 25, and Go varying with confining pressure; all other parameters maintain
their default values.
The undrained strength was characterized using the field testing programs and laboratory
tests presented by Abdellaziz et al. (2020). Fig. 3a shows a profile of undrained shear strength
(su) versus vertical effective stress (′vc) from the field exploration programs and laboratory tests.
The shear strengths were obtained from two CPT soundings, two field vane profiles, and five
monotonic triaxial simple shear (TxSS) tests. The undrained strengths increase linearly below a
′vc of approximately 25 kPa indicating a stress normalized behavior. The undrained strength
profile was fit with undrained strength ratios (su/ ′vc) of 0.4, 0.5, and 0.6 in Fig. 3a. The peak
su/ ′vc was characterized as 0.5 for these analyses. Figure 3b shows a profile of GMax versus ′vc
obtained from MASW and SCPTu data. GMax is approximately 18 MPa until a ′vc of 60 kPa
after which GMax linearly increases to approximately 65 MPa at a ′vc of approximately 95 kPa.
The calibration utilizes the profile indicated by the dashed line in Fig. 3b.

Figure 3: Profiles of (a) undrained shear strength (su), and (b) GMax versus vertical effective
stress (′vc)

PM4Silt is unable to directly simulate high peak strength ratios and large post-peak strength
loss because the high peak strength ratios require the model be close to critical state which does
not enable large strength loss to occur. A method was developed to address this limitation

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 590

through a numerical solution scheme wherein the critical state line (CSL) is dependent on the
shear strain. Fig. 4 presents a schematic that illustrates the influence of a strain dependent CSL
on the mobilized strength. The CSL shifts to the left in e-p′ space with increasing shear strain
which reduces the mean effective stress at critical state (p′cs). The CSL is shifted by reducing the
critical state line intercept at p′ = 1 kPa (e1). The reduced p′cs in turn reduces the deviator stress at
critical state (qcs) which reduces the mobilized peak strength.
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 4: Schematic illustrating influence of reducing e1 on the mobilized strength in (a) p′-
q space and (b) e-log(p′) space

Four calibrations were developed that varied the strain at the onset of strength loss and the
rate of strength loss. Figure 5 shows the stress-strain response for single element simulations of
DSS loading for the four calibrations. Calibration A has no post-peak strength loss and an su/ ′vc
= 0.5. Calibration B-D utilize the strength loss mechanism and have a St of approximately 12.
Calibration B initiated the strength loss mechanism at 5% shear strain (approximately the
mobilization of peak strength in Calibration A) and remolded over 10% shear strain. Calibration
C and D began to lose strength at 1% shear strain. Calibration C remolded over 50% shear strain
and calibration D remolded over 10% shear strain. Calibration C and D mobilized approximately
90% of the peak strength as the calibration was still hardening at the initiation of strength loss.

Figure 5: Stress-strain responses of single element monotonic DSS simulations of the four
calibrations

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 591

The cyclic behavior was iteratively calibrated to empirical expectations for sensitive clays
and to five strain-controlled cyclic TxSS tests at different shear strain amplitudes (0.3, 0.8, 1.0,
1.6, and 3.0%) at an ′vc = 0.6 atm by Abdellaziz et al. (2020). Fig. 6a shows the results of
simulations of stress ratio controlled cyclic DSS for Calibration A. Fig. 6b compares the
experimental TxSS simulations with strain-controlled cyclic DSS simulations for Calibration A.
The TxSS results indicate that the magnitude of strength loss increases with strain amplitude and
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

number of cycles. The calibration agreed relatively well with the experimental results at strain
levels of 1.6% and 3% shear strain and generally overpredicted strength loss at lower strain
levels.

Figure 6: Calibration A cyclic behavior: (a) stress ratio controlled DSS simulations, and (b)
strain controlled TxSS experimental data from Abdellaziz et al. (2020) and strain
controlled DSS simulations

The small strain stiffness degradation and damping of calibration A is compared with
experimental data in Fig. 7. Abdellaziz et al. (2020) developed G/Gmax (Fig. 7a) and damping
(Fig. 7b) curves from the results of 5 TxSS tests. The calibration reasonably approximates the
variation of shear modulus reduction with strain amplitude. The calibration underestimates
damping at small strains and overestimates damping at larger shear strains compared with the
experimental results as is commonly observed in numerical analyses.

Figure 7: Experimental data from Abdellaziz et al (2020) compared with single element
simulations of Calibration A for (a) modulus reduction and (b) damping

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 592

Crust and Till Characterization and Calibration

The crust was characterized using the data provided in Abdellaziz et al. (2020). The
undrained monotonic strength was characterized as having an su = 50 kPa with no post-peak
strength loss using the strength profile in Fig. 4. The small strain shear modulus was assumed to
be 18 MPa. The crust was modeled with PM4Silt with the parameters: su = 50,000 Pa, Go = 600,
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

hpo = 200; all other parameters retained their default values.


The till was assigned engineering parameters consistent with an SPT N value of 28 to 30 as
indicated by Abdellaziz et al. (2020). The till was modeled using PM4Sand with the parameters:
Dr = 81%, Go = 500 and hpo = 0.4; all other parameters retained their default values.

Non-linear Dynamic Analyses

Two-dimensional NDAs were performed using the finite difference program FLAC 8.1
(Itasca 2019) with the constitutive models PM4Silt (Boulanger and Ziotopoulou 2019) and
PM4Sand (Ziotopoulou and Boulanger 2016). The elements were 0.5 m tall x 0.5 m wide except
near the slope where the elements were modified to accommodate the slope. Simulations were
subjected to the 1988 Saguenay ground motion recorded at the Chicoutimi-Nord Station scaled
to a PGA = 0.1g as recommended by Abdellaziz et al. (2020). Fig. 8 illustrates the ground
motion time series (Fig. 8 left) and response spectra (Fig. 8 right) for the perpendicular
horizontal components (214 and 124). The predominant period for the 214 component was
around 0.2 to 0.4 seconds and the predominant period for the 124 component was approximately
0.2 seconds. The ground motions were applied as a shear stress time series to a compliant base.
Simulations were performed in 0.25 sec intervals with the shear strengths were modified based
on the elemental shear strain between each interval. Simulations were ended if the horizontal
displacement of the slope was greater than 1 m at the end of an interval.

Figure 8: Ground motion time histories (left) and response spectra (right) for the 124 and
214 components of the 1988 Saguenay earthquake recorded at the Chicoutimi-Nord station
scaled to a PGA = 0.1 g.

The initial stress conditions are illustrated in Fig. 9 which shows contours of the at-rest
horizontal earth pressure coefficient (Ko) and the initial static shear stress ratio (). The  values

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 593

generally range from 0.0 to 0.3 except directly beneath the dam face where  approaches 0.5 in
the soft clay and approximately 0.7 in the crust at the surface. Ko is generally between 0.5 and
0.7 in the regions away from the slope. Ko increases beneath the toe of the slope as the stress
ratios rotate and begin to approach the passive resistance stress ratio (Kp).
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 9: Contours of (a) Ko and (b)  at the end of consolidation prior to shaking

Analyses were performed using the four calibrations with the 214 component of the ground
motion. The simulation with Calibration A (non-strain softening) resulted in minimal
deformations (less than 1 cm) and the slope remained stable. The simulation with Calibration B
had similar results to Calibration A as the mobilized shear strains did not reach 5% and as such
no strength loss occurred. Calibrations C and D resulted in slope instability (horizontal
displacement > 1 m) about 17 seconds into the ground motion. These results indicated that the
stability of the slope was more dependent on the strain at the onset of strength loss than the rate
of strength loss.
Fig. 10 shows contours of shear strain capped at 15% (Fig. 10a) and horizontal displacement
(Fig. 10b) for the simulation with Calibration D at about 18 seconds into shaking, at which point
the slope was unstable/accelerating. The failure mobilized a circular failure mass which
predominantly failed through the base of the soft clay layer. The simulated failure daylighted
approximately 20 m from the toe of the slope. This was smaller than the failure observed in the
field however retrogressive landslides (which were not simulated herein) would likely further
extend the failure away from the toe of the slope.

DISCUSSION

The Saint Adelphe landslide provides a valuable case history to use to validate seismic slope
stability modeling approaches for soils with cyclic softening and strength loss. The slope was
marginally stable under static conditions due to the relatively steep slope and high water table
such that a relatively minor earthquake was capable of triggering the landslide. The approach
presented herein was able to reasonably predict the observed instability during the earthquake.
The numerical simulations have some degree of mesh dependency and further work is needed to
investigate the influence of strain concentrations on the seismic performance of earthen
structures.

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 594
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

Figure 10: Contours of (a) shear strain (capped at 15%) and (b) horizontal displacement at
the end of the simulation (failure) for Calibration D.

The constitutive modeling of the sensitive, high strength ratio, soft clay poses challenges.
The PM4Silt model (and other similar models) cannot model the large strength loss and high
peak undrained strength ratios because the mobilization of high strength ratios implies that the
model is nearly at critical state which does not allow for large post-peak strength loss. The
method herein for incorporating brittleness into PM4Silt provides a short-term solution for
simulating this behavior pending the development of more complete constitutive models for
structured sensitive clays. Further, characterizing the brittleness of sensitive clays is challenging
as evidenced by lab tests where the rate of post-peak softening does not appear to correlate with
strength loss observed in field studies. Further work is needed to characterize the rate of strength
loss in structured clays in the field.

CONCLUSION

NDAs of the 1988 Saint Adelphe landslide and a new method for incorporating brittleness
into analyses of clays with high peak strength ratios with the PM4Silt constitutive model were
presented. Brittleness was incorporated into the calibration using a strain dependent CSL, where
the accumulation of shear strain decreases p′cs and the associated qcs. The soft clay layer was
calibrated against available laboratory and field-testing data and was found to have high peak
strengths and low remolded strengths. The stability of the slope appeared to be more influenced
by the strain at the onset of strength loss than the rate of strength loss. The results of the NDAs
were shown to reasonably predict the observed instability during the earthquake.
Further work is needed to develop constitutive models that can directly capture large
sensitivity and high peak undrained strength ratios in nonlinear analyses. Nonetheless, the

© ASCE @seismicisolation
@seismicisolation
Geo-Congress 2023 GSP 340 595

procedures presented herein provided a short-term indirect methodology for evaluating the
dynamic behavior of structured, sensitive, high strength clays.

ACKNOWLEDGEMENTS

The work described herein was supported by the California Department of Water Resources
Downloaded from ascelibrary.org by Universidad Nacional Autonoma de Mexico on 01/23/24. Copyright ASCE. For personal use only; all rights reserved.

under Contract 4600009751. Any opinions, findings, or recommendations expressed herein are
those of the authors and do not necessarily represent this organization. The authors appreciate the
above support.

REFERENCES

Abdellaziz, M., Karray, M., Hussien, M. N., Delisle, M., Ledoux, C., Locat, P., Mompin, R., and
Chekired, M. (2020). “Experimental and numerical investigation of the Saint-Adelphe
landslide after the 1988 Saguenay earthquake.” Canadian Geotechnical Journal. 57: 1936-
1952.
Boulanger, R. W., and Ziotopolou, K. (2019). “A constitutive model for clays and plastic silts in
plane-strain earthquake engineering applications.” Soil Dynamics and Earthquake
Engineering, 1127(2019): 105832, 10.1016/j.soildyn.2019.105832.
Itasca. (2019). FLAC – Fast Lagrangian Analysis of Continua, Version 8.1, Itasca Consulting
Group, Inc, Minneapolis, Minnesota.
Kiernan, M., and Montgomer, K. (2020). “Numerical Simulations of the Fourth Avenue
Landslide Considering Cylic Softening.” Journal of Geotechnical and Geoenvironmental
Engineering. 146(1).
Oathes, T. J. (2022). Accounting for viscous effects in nonlinear analyses of strain softening
clays. Doctoral dissertation, University of California, Davis.
Ziotopoulou, K., and Boulanger, R. W. (2016). “Plasticity modeling of liquefaction effects under
sloping ground and irregular cyclic loading conditions.” Soil Dynamic and Earthquake
Engineering, 84(2016), 269-283, 10.1016/j.soildyn.2016.02.013.

© ASCE @seismicisolation
@seismicisolation

You might also like