You are on page 1of 110

COVER 4 COVER 1

3 BOLIVIAN
rd

3rd BOLIVIAN INTERNATIONAL CONFERENCE ON DEEP FOUNDATIONS


INTERNATIONAL
CONFERENCE ON
DEEP FOUNDATIONS

April 27 – 29, 2017


Santa Cruz de la Sierra, Bolivia

PROCEEDINGS
VOLUME 2
VOLUME 2

64428 ICDF Vol 2 Conference Cover_rb.indd 1 4/4/2017 10:14:47 AM


PROCEEDINGS
of the
3rd BOLIVIAN INTERNATIONAL
CONFERENCE ON DEEP FOUNDATIONS

April 27 – 29, 2017


Santa Cruz de la Sierra, Bolivia

VOLUME 2

The Bolivian Experimental Site for Testing Piles (B.E.S.T.)


Site Information

Edited by

Bengt H. Fellenius
K. Rainer Massarsch
Alessandro Mandolini
Mario Terceros Herrera
Design, execution, monitoring
and interpretation of deep foundation methods

ORGANIZING COMMITTEE

Conference Chairman:
Mario Terceros Herrera (Bolivia)

Conference Advisory Committee:


Bengt H. Fellenius (Canada)
Alessandro Mandolini (Italy)
K. Rainer Massarsch (Sweden)

Chairman of B.E.S.T. Prediction Event:


Bengt H. Fellenius (Canada)

© 2017, 3° C.F.P.B.

Copyright Information
Manuscripts are published according to an exclusive publication
agreement between the author and the conference organizer. Authors
retain copyright to their works.

Printed by Omnipress, Madison, WI USA


Table of Contents

Volume 2

Preface ........................................................................................................................v

Introduction ................................................................................................................1
Mandolini, Alessandro, Italy

Bolivian Experimental Site for Testing ...........................................................................................3


Fellenius, Bengt H., Canada, Terceros H. Mario, Bolivia
and Massarsch, K. Rainer, Sweden

Test Evaluation Reports ........................................................................................... 31

Boreholes and SPT-T .....................................................................................................................31


Decourt, Luciano, Brazil

Cone penetrometer soundings SCPTU Results..............................................................................35


Robertson, Peter K., Canada

Dilatometer tests ............................................................................................................................39


Marchetti, Diego, Italy

Pressuremeter tests .........................................................................................................................41


Frank, Roger, France

Interpretation of seismic tests ........................................................................................................47


Massarsch, K.Rainer, Sweden, Perez, A.M. Terceros and H. M. A Terceros., Bolivia

Interpretation of geotechnical parameters using in-situ data for B.E.S.T. ....................................69


Agaiby, Shehab S. and Mayne, Paul W., USA

ISSMGE TC10 – Seismic cone down-hole procedure to measure shear wave velocity:
A guideline .....................................................................................................................................91

3rd Bolivian International Conference on Deep Foundations—Volume 2 iii


iv 3rd Bolivian International Conference on Deep Foundations—Volume 2
PREFACE

The 3rd International Conference on Deep Foundations is held April 27 – 29, 2017 in Santa Cruz
de la Sierra, Bolivia. It follows two successful conferences held in 2013 and 2015. The conference
is organized with the support of INCOTEC SA in association with the Society of Engineers of
Bolivia, the Bolivian Society of Soil Mechanics and Geotechnical Engineering and the Chamber
of Construction of Santa Cruz. It is held at the UPSA Campus (Universidad Privada de Santa
Cruz), the main private university of the city and arranged with the support of the International
Society of Soil Mechanics and Geotechnical Engineering (ISSMGE), Technical Committee 212,
“Deep Foundations”.

The principal objective of the conference is to bring together local engineers and international
experts in order to facilitate the exchange of experience and to introduce to the region new design
concepts, methods and equipment for the application to deep foundations. The conference program
is composed of invited lectures, discussions, a field demonstration, and a pile testing prediction
event where international experts have been invited to predict the load-movement response of piles
in static loading carried out prior to the conference.

During the first two days of the conference, speakers of international repute have been invited to
present papers on specific topics, covering different aspects of deep foundations. The third day of
the conference is devoted to the presentation and discussion of tests a comprehensive pile testing
program. The Bolivian Experimental Site for Testing Piles (B.E.S.T.) was adopted by ISSMGE
TC 212 as a reference site for investigations on piles and pile groups. B.E.S.T. offers a unique
possibility to enhance the understanding of the performance of different pile types and pile groups
when subjected to load. The geotechnical conditions at the B.E.S.T. site have been documented
by detailed investigations, using state-of-the art testing and interpretation methods. The results of
the field testing programme, including interpretation of in-situ methods and results of the pile
loading tests will be presented during the third day of the conference.

Volume 1 of the proceedings comprises the papers presented at the conference. All papers have
been reviewed by at least two members of the Review Committee. The dedicated work by the
reviewers and their valuable contributions is gratefully acknowledged.

Volume 2 contains a description of the geological setting and the results of comprehensive
geotechnical investigations carried out at the B.E.S.T. site. It is the intention of the Conference
Organizers to make available all data from the B.E.S.T. site investigations and pile tests in digital
format at the conference web platform for use in future investigations, in cooperation with
ISSMGE TC 212.

Volume 3 includes a description of the test piles and the loading test programme. The predictions
as well as a presentation of test results will be published in a Volume 3 after the conference.

3rd Bolivian International Conference on Deep Foundations—Volume 2 v


vi 3rd Bolivian International Conference on Deep Foundations—Volume 2
INTRODUCTION

Mandolini, A,(1) (Italy), Chairman of ISSMGE TC212

Università degli Studi della Campania “Luigi Vanvitelli”, Italy


<alessandro.mandolini@unina2.it>

In 2013 and 2015, Mario H. Terceros of INCOTEC, a foundation engineering company


headquartered in Santa Cruz, Bolivia, organized the 1st and 2nd Bolivian International Conference
on Deep Foundation. The conferences had major success due to the technical and scientific
contribution of many specialists from all over the world. An important aspect of these conferences
was the prediction events of pile loading tests which attracted contributions from piling experts
world-wide.
At the 2nd conference, I had the great honor to be invited by the organizers to deliver a lecture
on “Sustainable Piling Engineering”. In addition to the technical events, visits to historic sites were
arranged, making it possible to enjoy that beautiful country that Bolivia is, as well as the excellent
hospitality by the Bolivian conference organizers and, in particular, Mario and his team of
competent engineers.
During my stay in Santa Cruz, ISSMGE made known the results of the election for the
Chairmanship of TC212 “Deep Foundation”: in some way unexpected, I learned my appointment
as Chairman for the period 2015-2019. The resulting enthusiasm for this appointment, combined
with Mario’s willingness to invest in applied research, formed the basis for some preliminary
discussion about the opportunity to set up a local experimental site, in the following named
“Bolivian Experiment Site for Testing” (B.E.S.T.), situated 24 kilometers North-East of Santa
Cruz de la Sierra.
The fundamental idea of establishing B.E.S.T was, and is, that of having available a site where
subsoil conditions are well known (geology, stratigraphy, physical and mechanical soil properties,
groundwater regime, etc.) and where different geotechnical structures can be installed and tested;
needless to say, the first option was that of concentrating efforts to install different pile types in
order to compare their performance when loaded (B.E.S.T. Piles).
The idea was presented at the TC212 meeting held in Edinburgh at the occasion of the XVI
ECSMGE in September 2015. During the discussion, it became apparent that this initiative would
fit the ISSMGE objectives, in particular: to disseminate and to develop knowledge and practice;
to interact with industry and overlapping organizations working with pile foundations.
Consequently, B.E.S.T. for piles was included in the TC212 Terms of Reference and, on 30th
January 2016, the initiative was announced on the ISSMGE website.
A small organizing group, consisting of Bengt H. Fellenius (Coordinator), Alessandro
Mandolini, K. Rainer Massarsch, and Mario H. Terceros, was given the task to organize the pile
testing programme addressing static and dynamic pile tests in connection with the 3rd conference.
Possibly, Mario underestimated that “hunger comes during eating” but, thanks to his
enthusiasm, he made possible to install 22 piles (bored, self-boring, CFA, and full displacement)
having same length (9.5 m) and diameters ranging from 150 mm to 620 mm. All the piles were
instrumented along the shaft in order to obtain information about load sharing between base and
shaft. Six single piles were equipped with enlarged pile toe. In addition, a piled raft containing 13
full displacement piles, were tested.

3rd Bolivian International Conference on Deep Foundations—Volume 2 1


It was clear from the beginning of this adventure that such a so massive research programme
would give the opportunity to significantly improve knowledge about pile behavior, but also to
spread this unique information to the technical and scientific community.
The organizing committee, consisting of a small group of enthusiasts came to the conclusion
to organize a “Class A” International Prediction Event; Mario Terceros offered his availability to
host a full day conference within the framework of the 3rd Bolivian International Conference on
Deep Foundation held in Santa Cruz in April 2017.
Outstanding experts from all over the world were asked to contribute to B.E.S.T presentation
of the results. All of them are gratefully acknowledged.
On behalf of TC212, I hope that this exciting experience will help all of us to achieve an
ambitious target, that of inspiring design criteria aimed to increase the safety with lower factor of
safety!!!

2 3rd Bolivian International Conference on Deep Foundations—Volume 2


BOLIVIAN EXPERIMENTAL SITE FOR TESTING

Presentation of Field Testing Programme


Fellenius, B.H.(1), Terceros H. M.(2) and Massarsch. K.R.(3)
(1)
Consulting Engineer, Sidney, BC, Canada <bengt@fellenius.net>
(2)
Incotec S.A., Santa Cruz de la Sierra, Bolivia <math@incotec.cc>
(3)
Geo Risk & Vibrations Scandinavia AB, Bromma, Sweden <rainer.massarsch@georisk.se>

1. BACKGROUND
The purpose of the “Bolivian Experimental Site for Testing” (B.E.S.T) is to provide well-
documented, comprehensive geotechnical information from a site where different types of pile
tests have been performed. The results from the field investigations and the full-scale pile loading
tests are intended to augment the current state-of-the-art of pile design. The results of the field
studies are being presented in connection with the 3rd International Conference on Deep
Foundations, Bolivia (C.F.P.B.), held in Santa Cruz from April 27 through 30, 2017.
Pile design methods have advanced significantly. However, they still rely greatly on empirical
correlations. In spite of significant recent improvements made in identifying the complex actions
occurring when a pile is installed or constructed, as well as the process of long-term load-transfer,
the interaction between structure, piles, and soil to consider in the design of piled foundations,
needs much further study. Currently, new types of piles, installation and construction methods
have emerged. Such new pile types are, for example, drilled displacement piles, full displacement
piles, Expander Body piles, Toe Box piles, helical piles, injected micro piles, etc. Although these
new pile types, potentially, offer significant savings and improved quality, design methods are still
based on simplified concepts. Better knowledge regarding their validity in relation to existing
practice is needed by the deep foundation industry.
The current state-of-the-art of piled foundation design relies on empirical correlations and
emphasizes capacity, that is, ultimate shaft and toe resistances. An important development of piling
technology has been the possibility to monitor and control the installation process. This
information can be used to determine the required depth of installation and to estimate, based on
empirical correlations, the bearing capacity of piles after installation.
Settlement analysis is rarely included and less understood. The design analysis is usually
limited to the response to load applied to single piles and, while verification tests on single piles
are common, tests on pile groups are extremely rare—only a handful reports are available and they
are frequently based on results of model studies. Therefore, the understanding of pile group
response in piled foundation design is limited.
Moreover, current design practice relies primarily on soil description from samples obtained
from boreholes (BH) and Standard Penetration tests (SPT) N-indices and, in some areas, also Cone
Penetration Tests with pore water pressure measurement (CPTU). The use of other in-situ tests,
such as dilatometer (DLT), pressuremeter (PMT), shear wave velocity measurements (also in
combination with CPT and DMT) or seismic surface wave measurements (MASW) is still rare.

3rd Bolivian International Conference on Deep Foundations—Volume 2 3


2. PILE TEST PROGRAMME
The pile test programme reported herein aims to determine the load movement relationship of the
pile head and pile toe for the different pile types, which includes determining the axial load
distribution in the piles for different loads applied to the pile head. Loading tests are performed on
individual piles and pile groups. In addition, pile integrity tests and dynamic tests are being carried
out. Full-scale instrumented pile tests to study load transfer and settlement will be performed on
single piles and on a pile group. Three single piles and the pile group will be used for a prediction
event. A few piles are installed with intentional defects to serve as a base for assessing the
application and limitations of integrity testing methods. The detailed test programme and
scheduling is presented in subsequent sections.
The results of the pile loading tests and a compilation of pile predictions submitted prior to
the loading tests is presented in Volume 3 of the proceedings.

3. GEOLOGY AND SOIL PROFILE


The city of Santa Cruz de la Sierra, lies in the southern part of the Amazonas and the B.E.S.T. site
is situated in the Warnes municipality, 24 kilometers North-east of Santa Cruz de la Sierra (see
Figure 1). The geology of the area is characterized by a Paleozoic (250 million years old)
sedimentary basin. The soils of interest to civil engineering construction, the surficial soils, are
quaternary with the dominant minerals being calcite, silica, and feldspar. The main agent is the
Piray River and its tributaries, which past meandering over the area has resulted in a sedimentation-
erosion-sedimentation process and a geological profile dominated by fine to medium sands with
intermittent layers of clay or clayey sand. The geology in and around the city, with intermittent
layers of compressible soils, is such that even light buildings need to be supported on piles.

Fig. 1. Location of the city of Santa Cruz and the B.E.S.T. site in reference to Santa Cruz.

In summary, the upper about 10 to 20 m part of the profile consists of normally consolidated
layers of clays, silts, sands, in various combination and thickness. The upper about 5 to 6 m consists
of loose silt and sand. Hereunder lies a 6 to 7 m layer of compact silt and sand. At about 11 m
depth lies an about 1 m thick layer of soft silty clay followed by an about 1 m thick layer of compact

4 3rd Bolivian International Conference on Deep Foundations—Volume 2


sand. Below about 12 m depth, the profile alternates between about 2 m thick layers of compact to
dense silty sand and about 2 m thick layers of loose sand. The groundwater table at the site ranges
seasonally between the ground surface and about 0.5 m depth.

4. SOIL EXPLORATION
The B.E.S.T. site is about 40 m wide and 100 m long. The geotechnical conditions of the site have
been investigated using conventional in-situ and laboratory methods. At each single pile location
and at the location of two piles of the group (E-piles), the following in-situ tests have been
performed:

• SPT (standard penetration test) and SPT-T (SPT plus torque measurement).
Dynamic measurements are made to determine the transferred energy ratio
(ETR)
• SCPTU (seismic piezocone penetration test with pore water pressure
measurement)
• SDMT (seismic dilatometer test)
• PMT (pressuremeter tests)
• SASW and REMI geophysical tests

Each borehole and field test is identified with the letter of its designated test pile and located at
0.80-m radial distance from the test pile. The first four in-situ tests are placed in the corners of a
square inscribed in an octagon with the designated pile in its center and its sides parallel with the
line between the test piles; Figure 2 shows the locations and denotations for in-situ tests around
Pile A-1. Any additional in-situ tests will be placed in the opposite corners of the octagon.

Fig. 2. Example of layout of tests performed adjacent to Pile A-1.

When possible, water content and grain size distribution were determined for samples. The
samples were classified for soil type per the Unified Soil Classification System (USCS). For
cohesive samples, consistency limits (LP and LL) were also determined. Section 9 shows the
compilations from all site borehole data, and SPT and CPT tests at the B.E.S.T. site.

3rd Bolivian International Conference on Deep Foundations—Volume 2 5


5. SINGLE PILE TESTS
5.1 Placement and Construction Details

Different types of single piles and one pile group will be constructed and tested at the site. Figure
5.1 shows the pile locations and Table 5.1 shows a summary of the different piles and tests. The
head-down tests will be using four reaction piles placed in a square configuration with a 5 m center
to center distance with the test pile in the center (the distance between test pile and anchor pile is
3.5 m). The piles intended for static testing will be installed to a depth of 9.5 m, i.e., about 1.5 m
above the about 1 m thick soft clay layer. (The length of the piles intended for integrity testing
demonstration will be disclosed only after the demonstrations are completed).
Locations of Pile Tests, Boreholes, and In-situ Tests
South to North (m)

6.0 A1 A2 A3 B1 B2 C1 C2 F1 F2 G1 D1 D2 E1 E2-E13 H1
3.0
0.0
0 10 20 30 40 50 60 70 80 90
East to West (m)
Pile Group, Piles E2 - E14
Fig. 3. Test pile locations.

All piles subjected to static testing are instrumented with strain-gages for measurement of
axial load distribution. Piles A1, B1 and B2, C1 and C2, D1 and D2, and all E-piles will have the
reinforcement cage equipped with an Expander-Body (EB) to ensure a toe resistance larger than
the shaft resistance. The EB is constructed by expanding a folded steel cylinder by injecting grout.
The 9.5-m installation depth is the depth of the EB end before inflation. A bidirectional cell (BD)
is placed above the EB in all EB-equipped piles with the BD bottom plate at 8.3-m depth, as shown
in Figure 4. The EB width prior to expansion is 160, 140, 120, and 110 mm in Piles A3, B2, C1,
D1, and E-piles, respectively. After installation and concreting the pile shaft, the EB is pressure-
grouted expanding the diameter to 815, 600, 500, and 300 mm for Piles A3, B2, C1, D1, and E-
piles, respectively. Pressure-grouting the soil below the EB completes the installation. Piles B2,
C2, E1, and the E-pile group will be included in a prediction event.

Pile Head
Ground surface
Pile toe depth = 9.5 m

Bidirectional Cell

Inflated EBI
inflation
1.20 m
Before

Grouted zone
After
inflation
Fig. 4. Sketch showing the EB and BD arrangement (not to scale).

6 3rd Bolivian International Conference on Deep Foundations—Volume 2


The pile construction procedures are as follows: The cylinder strength of concrete and mortar is
designed to be 30 MPa (H4,300 psi). The reinforcement cage consists of six 16-mm bars in Piles
A3, B2, and C1 and four 12-mm bars in Piles E. All reinforcement cages have a 6-mm spiral with
a 250-mm pitch. All test piles are intended to be constructed to 9.5 m depth. All reinforcement
cages are instrumented with strain-gages and bidirectional cell at the bottom end.

A. Three 620-mm diameter, bored piles, A1, A2, and A2, constructed using a slurry. Once the
final depth is reached, the 500-mm reinforcement cage is placed in the pile. Thereafter, the
concrete is tremied (pumped into the shaft starting at the toe of the pile). Pile A1 will have
an EB800 placed below the BD. Pile A2 will have a "Toe Box" (TB) placed below the BD,
which is a 500-mm diameter and 300 mm high steel telescopic device installed with the
reinforcing cage. After the shaft concrete is hardened, grout is pumped into the TB
expanding the height of the box and compressing the soil below the pile toe and introducing
a compressive strain in the pile shaft. The TB will be equipped with a "Shaft Box" (SB),
which is a folded steel belt welded to the outside of the TB. When the TB grout has
hardened, grout is pumped into the SB expanding the width of the TB-SB to about 800 mm
diameter compressing the soil around the TB. Pile A3 has neither EB nor TB. Pile A3 is a
part of the prediction event.
B. Two 450-mm diameter continuous flight auger (CFA), partial displacement piles, B1 and
B2. The central stem of the auger is 250 mm of O.D. Mortar is used instead of concrete in
order to allow the subsequent installing the reinforcement cage. Pile B-1 has a bidirectional
cell (BD) and an Expander Body (EB800) placed below the BD. Pile B2 are straight (have
no EB) and are a part of the prediction event.
C. Two 450-mm diameter Full Displacement Piles (FDP) with "lost bit", C1 and C2. The
equipment is illustrated in Figure 5 and consists of a 450-mm O.D. displacement body
(pipe) with a 800 mm long bulb attached to a 1.15 m long auger with a 350-mm diameter.
The auger rotation pulls down the displacement body. The drill equipment is a Bauer
BG18PL with a 180-kMm maximum torque. The auger has a short conical tip that is left
in the hole upon completion (“lost bit”). Pile C1 will have an Expander Body (EB600) at
the toe and Pile C2 is straight. After placing the reinforcement cage in the casing, concrete
is pumped into the shaft starting at the toe of the pile gradually withdrawing and the casing.
The "lost bit" remains in the ground. Pile C2 is a part of the prediction event.

3rd Bolivian International Conference on Deep Foundations—Volume 2 7


C-PILES E-PILES

260 mm

220 mm

450 mm
800 mm

300 mm
800 mm

350 mm 1,150 mm

220 mm 250 mm

Fig. 5. Sketch showing equipment and geometry of the full displacement pile (not to scale).

D. Two 150-mm diameter self-boring micropiles, D1 and D2, with a 75-mm diameter drilling
pipe and a cutting tool at the pipe end. Fluid grout is injected as the pile penetrates the soil.
Pile D1 will have an Expander Body (EB600) at the toe and Pile D2 will be straight. The
reinforcement cage consists of six 12-mm bars inside a 6-mm spiral with a 250-mm pitch.
The drilling pipe remains in the pile. A solid 32-mm diameter bar will be placed inside the
Pile-D1 drilling pipe to increase the axial stiffness of the shaft.
E. Fourteen 300-mm diameter FDP piles, E1 through E14. The equipment is illustrated in
Figure 5.3 and consists of a 220-mm O.D. displacement body (pipe) with a 800 mm long
bulb attached to a 0.25 m long auger with a 220-mm diameter (no lost bit). The auger
rotation pulls down the displacement body. The drill equipment is a Bauer BG18PL with
a 180-kNm maximum torque. All piles to have an EB300 at the toe and a BD above the
EB. After placing the reinforcement cage in the casing, concrete is pumped into the shaft
starting at the toe of the pile gradually withdrawing the casing. Pile E1 test will be a part
of the prediction event.
F. Three bored piles with different diameter, 450, 600, and 1,200 mm, respectively, F1 - F3,
constructed using a slurry. Once the final depth is reached, the reinforcement cage with
attached gages and bidirectional cell (BD) is placed in the shafts of F1 and F2, but not in
F1. Thereafter, 30-MPa strength concrete is tremied (pumped into the shaft starting at the
toe of the pile. Static loading tests will be carried out on Piles F1 and F2. Pile F3 is intended
for integrity testing, only.
G. 300 mm diameter long helical pile, G1.

8 3rd Bolivian International Conference on Deep Foundations—Volume 2


H. Spare test locations for use for potentially needed back-up piles.
I. Five piles denoted DC1200-1 through DC1200-5 will be installed with intentional, but
undisclosed, defect(s) and various integrity testing methods will be used for comparing the
results of the tests to the actual defects.

The results of the tests on Pile Types A through D are intended to be used for comparing the
response of same type piles with and without toe enhancement (EB, TB, and SB) and to the other
pile types in terms of stiffness, shaft and toe responses, and ultimate resistance. The results of the
test on Pile E1 will serve as reference to the results of the tests on the E-piles. Phase 1 is a
bidirectional test. The BD test on the E-pile group will apply equal force to the piles, while
measuring the upward and downward movements for each pile at the cell level and the pile head
(the movements will differ depending on the position in the group). The second, Phase 2, is a head-
down test. For the E-pile group, the piles are encased in a rigid pile cap cast on the ground. Thus,
the head-down test will involve equal movement of all pile heads and measuring the resulting pile
forces by means of a load cell placed on the pile heads immediately below the pile cap (before
casting the cap).

5.2 Pile Instrumentation


The pile head movement is measured against two reference beams supported on the ground at
a 2.5 m distance from the test pile and placed parallel to the pile center line.
Each BD-equipped pile will have two pairs of telltales placed diametrically opposed. One pair
will measure the movement between the pile head and the bottom of BD plate and one will measure
between the pile head and the upper BD plate. Subtracting one from the other will give the opening
of the BD cell.

3rd Bolivian International Conference on Deep Foundations—Volume 2 9


TABLE 1. Pile and Test Summary.

10 3rd Bolivian International Conference on Deep Foundations—Volume 2


Pairs of sister-bar strain-gages manufactured by Geovan Geotechnical Instruments
(www.Geovamn.com) are placed in each of the single test piles at 2.0, 5.0, and 7.5 m depths.
Toe accelerometers are placed in piles Type A, B, C, F, and G in order to indicate the toe
movement during the dynamic testing performed after completion of the static loading test
programme.
Pipes for Cross Hole Test are provided in Piles A, B, and C
Thermal Wires are installed in Piles A, C, and E.
The BD pressure (load) and the upward and downward movement of the pile at the BD level
and at the pile head will be measured using two pairs of telltales.
At the E-group piles, the soil is instrumented to measure movement between the pile and the
soil and soil strain during Phases 1 and 2. They are placed at 0.5 m depth below the slab and at the
pile toe level using a soil anchor system at two points: mid-way between Piles E5 and E6 and Piles
E10, and E11. Two earth-stress cells are also placed to measure contact stress in Phase 2 of the E-
group test: below the pile cap at mid-way between Piles E5 and E10 and Piles E6, and E11
One soil anchor is also installed near the single pile, Pile E1 before the pile is installed and
placed at the pile toe level at an about 0.35-m distance from the pile surface.

5.3 Pile Test Methods and Procedures


The test procedure is according to the "quick method" consisting of a series of equal load
increments applied at 15-minute time intervals and held constant (electrically operating, automatic
pressure-holding pump is required) during each interval until excessive movements have
developed, whereupon the pile is unloaded in five or six about equal decrements, each maintained
for 5 minutes.
The load increments are about 5 % of estimated maximum load. N.B., the test will continue
until reaching either the pile plunges or the limit of available force, or the limit of expansion of the
BD cell or of the jack is reached. The total test duration is, therefore, about 5+ hours. All gage
readings will be recorded in a common data logger (multi-channel) and 30-s scan intervals. (All
tests must have a data logger capable to record all data simultaneously in order to ensure that the
data logger will scan and record all gages to a common file and time stamp).

6. PILE GROUP TESTS


The pile group is made up of 13 Type E piles (Piles E2-E14) installed by as FDP piles at a c/c
of 2.5b (0.6 m) configured as shown in Figure 6. The soil within the E-piles is instrumented with
gages measuring strain (soil shortening over a short distance) underneath the pile cap and at the
pile toe level.
After the completion of the BD tests, Phase 1, the single pile and the group piles are subjected
to a head-down test with, Phase 2, open BD and, Phase 3, closed BD.
Phase 2 will start after the completion of Phase 1 by placing a load cell on each pile head,
then, a 0.2 m thick compacted sand layer is placed around the piles and a rigid reinforced concrete
slab connecting the piles is cast on the ground encompassing all piles. The slab is cast directly on
to the sand layer so as to ensure a contact between the group slab and the soil. Before casting the
slab, a 200 mm thick sand layer is placed on the ground and compacted. The slab is loaded by
means of weights or by jacking it against a loaded platform, engaging all 13 group piles. The BD
units are open (free-draining) in the Phase-2 test.
Potentially, a Phase 3 will be performed as a repeat of Phase 1 with the pile rigid cap (platform)
remaining but unloaded.

3rd Bolivian International Conference on Deep Foundations—Volume 2 11


2.54 m
E2 E3 E4

E5 E6

E7 E8 E9

E10 E11

E12 E13 E14

Fig. 6. B.E.S.T. Pile group configuration (E2 – E14).

7. PREDICTION EVENT
Single Piles, A3, B2, C2, and E1, and group Piles E2 - E14 are part of a prediction event addressing
the pile response in terms of load-movement. The event will be reported in Volume 3 of the
conference proceedings.

8. INTEGRITY TESTS and DEMONSTRATION


An integrity testing field demonstration is held on Friday April 28 immediately before the
conference. The piles intended for the demonstration will have been supplied with intentional
flaws of different types. The design of the flaws has been discussed with the participating
commercial laboratories specializing in pile integrity test, but how the flaws are arranged in the
test pile and, indeed, if an intentional flaw is included in a specific pile, is not disclosed to the
participants in the demonstration. All "integrity" piles will include thermal wires. The Integrity
Demonstration is reported separately. The pile types included are:

• Pile DC1200-1, 1,200-mm diameter bored pile with retrievable casing


• Piles DC620-2 and DC620-3, 620-mm diameter bored pile with retrievable casing
• Piles CFA450-1 and CFA 450-2, 450-mm FDP piles

9. COMPILATION OF BOREHOLE INFORMATION AND IN-SITU TESTS


9.1. Boreholes
Eight boreholes were drilled with SPT split-spoon sampling. Three (BH-B2, BH-F1, BH-G1)
to 9.5 m depth and five (BH-A3, BH-C1, BH-D1, BH-E1, and BH E2)to 25 m depth. The
distributions of grain size, water content, plastic and liquid limits, and SPT N-indices (with qt-
diagram) have been compiled in a single set of diagrams from each borehole as shown below.
Electronic format information is available for downloading.
The SPT hammer weight was 62.5 kg and the free-fall height was 760 mm according to the
ISSMGE reference procedure. Nominal hammer energy is 466 J. Dynamic tests were carried out

12 3rd Bolivian International Conference on Deep Foundations—Volume 2


at BH-B2 at 6 through 09.5 m depth. The transferred energy ratio (ETR) for the test are shown in
Figure 7.1. The ETR mean value is 44 %.

25
MEAN = 44 %
STD Dev. = 6.9 %
No. of blows = 115
F 20
r
e
q
u 15
e
n
c
y 10

(%)

0
10 20 30 40 50 60
Energy Transfer Ratio, ETR (%)

Fig. 7. Histogram of transferred energy at BH-B2, depths 6 m through 9.5 m.

Water content was determined on a sample from all depths. Plastic and liquid limits were
determined on samples where visual inspection suggested cohesive condition. The water content
values and Atterberg limits together with the fines contents show most of these samples to be silt
more than clay.
The "Soil Type Fraction" indicates the amount of fines in percent of total sample (Sieve #200).
No gravel was found in the samples.

9.2. Summary of Borehole Data


The results of the laboratory and SPT investigations of boreholes: BH-A3, BH-B2, BH-C1, BH-
D1, BH-E1, BH-E2, BH-F1 and BH-G1 are shown Figures 8 through 15.
The as-recorded SPT N-indices are shown in bars and the CPTU qt-curve is overlaid
the N-index diagram. Note, as the boreholes and the CPTU soundings (15 in total) are about 1 to
2 m apart, the soil boundary depths indicated by the N-index bar and water content curve do not
always agree with boundary depth suggested by the qt-curve.
All the five 25-m deep boreholes, show an about 1 m thick clay and silt layer at about 11 m
depth (12 m in BH-D1) immediately above a same thickness sand layer.

3rd Bolivian International Conference on Deep Foundations—Volume 2 13


BH-A3 SPT N-INDICES
WATER CONTENT (%) SOIL TYPE FRACTIONS (%)
(blows/0.3 m)
0 5 10 15 20 25 30 35 40 45 50 0 20 40 60 80 100 0 10 20 30
0 0 0
GW SAND
2 2 2
wn
CLAY and SILT
4 4 4

6 6 6

8 8 8
DEPTH (m)

10 10 10

12 12 CLAY and SILT 12


SAND
14 14 14

16 16 16
CLAY and
SILT SAND
18 18 18

20 20 20

22 22 22

Fig. 8. Borehole data and SPT N-values, BH-A3.

BH-B2 SOIL TYPE FRACTIONS (%) SPT N-INDICES


WATER CONTENT (%)
(blows/0.3 m)
0 5 10 15 20 25 30 35 40 45 50 0 20 40 60 80 100 0 10 20 30
0 0 0
GW CLAY SAND
2 2 and SILT 2
PL wn LL
4 4 4

6 6 6

8 8 8
DEPTH (m)

10 10 10

12 12 12

14 14 14

16 16 16

18 18 18

20 20 20

22 22 22

Fig. 9. Borehole data and SPT N-values, BH-B2.

14 3rd Bolivian International Conference on Deep Foundations—Volume 2


SPT N-INDICES
BH-C1 WATER CONTENT (%) SOIL TYPE FRACTIONS (%) (blows/0.3 m)
0 5 10 15 20 25 30 35 40 45 50 0 20 40 60 80 100 0 10 20 30
0 0 0
GW
2 2 2
SAND
PL wn LL CLAY and SILT
4 4 4

6 6 Loose Gravel 6

8 8 8
SAND
DEPTH (m)

10 10 10

12 12 Clay 12
Silty Sand
14 14 14
CLAY
SAND
16 16 and 16
SILT
18 18 18

20 20 20

22 22 22

Fig. 10. Borehole data and SPT N-values, BH-C1.

BH-D1 SPT N-INDICES


WATER CONTENT (%) SOIL TYPE FRACTIONS (%)
(blows/0.3 m)
0 5 10 15 20 25 30 35 40 45 50 0 20 40 60 80 100 0 10 20 30
0 0 0
GW wn CLAY
2 2 2
and SAND
4 SILT 4
4

6 6 6

8 8 8
SAND
DEPTH (m)

10 10 10

12 12 Clay 12
CLAY and SILT
14 14 SAND 14

16 16 16
CLAY
18 18 and 18
SILT
20 20 20

22 22 22

Fig. 11. Borehole data and SPT N-values, BH-D1.

3rd Bolivian International Conference on Deep Foundations—Volume 2 15


BH-E1 SPT N-INDICES
WATER CONTENT (%) SOIL TYPE FRACTIONS (%) (blows/0.3 m)
0 5 10 15 20 25 30 35 40 45 50 0 20 40 60 80 100 0 10 20 30
0 0 0
GW
2 2 2
PL wn LL SAND
CLAY and SILT
4 4 4

6 6 6
Sand
8 8 8
SAND
DEPTH (m)

10 10 10

12 12 Clay and Silt 12


Sand
14 14 14
CLAY
16 16 16
and
SAND
SILT
18 18 18

20 20 20

22 22 22

Fig. 12. Borehole data and SPT N-values, BH-E1.

BH-E2 SPT N-INDICES


WATER CONTENT (%) SOIL TYPE FRACTIONS (%)
(blows/0.3 m)
0 5 10 15 20 25 30 35 40 45 50 0 20 40 60 80 100 0 10 20 30
0 0 0
GW
2 2 2
SAND
PL wn LL CLAY and SILT
4 4 4

6 6 6

8 8 8
SAND
DEPTH (m)

10 10 10

12 12 Clay and silt 12


Sand
14 14 14
CLAY
SAND
16 and 16
16
SILT
18 18 18

20 20 20

22 22 22

Fig. 13. Borehole data and SPT N-values, BH-E2.

16 3rd Bolivian International Conference on Deep Foundations—Volume 2


BH-F1 SPT N-INDICES
SOIL TYPE FRACTIONS (%)
WATER CONTENT (%) (blows/0.3 m)
0 5 10 15 20 25 30 35 40 45 50 0 20 40 60 80 100 0 10 20 30
0 0 0
GW SAND
2 2 CLAY and SILT
2

4 4 4
wn
6 6 6
CLAY and SAND
8 8 SILT 8
DEPTH (m)

10 10 10

12 12 12

14 14 14

16 16 16

18 18 18

20 20 20

22 22 22

Fig. 14. Borehole data and SPT N-values, BH-F1.

BH-G1 SPT N-INDICES


WATER CONTENT (%) SOIL TYPE FRACTIONS (%)
(blows/0.3 m)
0 5 10 15 20 25 30 35 40 45 50 0 20 40 60 80 100 0 10 20 30
0 0 0
GW PL wn LL
2 2 CLAY and SILT 2

4 4 4
CLAY SAND
6 and SILT 6
6

8 8 8
DEPTH (m)

10 10 10

12 12 12

14 14 14

16 16 16

18 18 18

20 20 20

22 22 22

Fig. 15. Borehole data and SPT N-values, BH-G1.

9.2 Cone Penetrometer Soundings


The cone penetrometer soundings with pore water pressure measurements are presented in the
following diagrams, Figures 16 - 30. For text files with the results of the cone penetrometer tests,
see the conference web site. Figure 16 is a compilation of all cone stress measurements, qt,

3rd Bolivian International Conference on Deep Foundations—Volume 2 17


corrected for pore water pressure. Note that the line indicating the neutral pore pressure distribution
is only a fit to the U2-pressure and does not necessarily indicate the depth to the groundwater table.

Cone Stress, qt (MPa)


0 5 10 15
0
A1
A2
A3
5 B1
B2
C1
D1
10
DEPTH (m)

D2
E1
E2
15 F1
F2
F3
G1
20
G2

25

Fig. 16. CPT qt records from 15 tests.

CPTU A1
Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
Cone Stress, qt (MPa)
0 25 50 75 100 0 200 400 600 800 1,000 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10
DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 17. Results of CPTU A1.

18 3rd Bolivian International Conference on Deep Foundations—Volume 2


CPTU A2
Cone Stress, qt (MPa) Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
0 25 50 75 100 0 100 200 300 400 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10

DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 18. Results of CPTU A2.

CPTU A3
Cone Stress, qt (MPa) Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
0 25 50 75 100 0 100 200 300 400 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10
DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 19. Results of CPTU A3.

3rd Bolivian International Conference on Deep Foundations—Volume 2 19


CPTU B1
Cone Stress, qt (MPa) Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
0 5 10 15 0 50 100 150 200 0 200 400 600 800 1,000 0 1 2 3 4 5
0 0 0 0

5 5 5 5

10 10 10 10

DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 20. Results of CPTU B1.

CPTU B2
Cone Stress, qt (MPa) Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
0 5 10 15 0 25 50 75 100 0 100 200 300 400 0 1 2 3 4 5
0 0 0 0

5 5 5 5

10 10 10 10
DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 21. Results of CPTU B2.

20 3rd Bolivian International Conference on Deep Foundations—Volume 2


CPTU C1
Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
Cone Stress, qt (MPa)
0 25 50 75 100 0 200 400 600 800 1,000 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10

DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 22. Results of CPTU C1.

CPTU D1
Cone Stress, qt (MPa) Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
0 25 50 75 100 0 100 200 300 400 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10
DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 23. Results of CPTU D1.

3rd Bolivian International Conference on Deep Foundations—Volume 2 21


CPTU D2
Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
Cone Stress, qt (MPa)
0 25 50 75 100 0 200 400 600 800 1,000 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10

DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 24. Results of CPTU D2.

CPTU E1
Cone Stress, qt (MPa) Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
0 25 50 75 100 0 100 200 300 400 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10
DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 25. Results of CPTU E1.

22 3rd Bolivian International Conference on Deep Foundations—Volume 2


CPTU E2
Cone Stress, qt (MPa) Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
0 25 50 75 100 0 200 400 600 800 1,000 0 1 2 3 4 5
0 5 10 15
0 0 0
0

2 2 2
2

4 4 4 4

6 6 6 6

8 8 8 8

DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

10 10 10 10

12 12 12 12

14 14 14 14

16 16 16 16

18 18 18 18

20 20 20 20

22 22 22 22

Fig. 26. Results of CPTU E2.

CPTU F1
Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
Cone Stress, qt (MPa)
0 25 50 75 100 0 200 400 600 800 1,000 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10
DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 27. Results of CPTU F1.

3rd Bolivian International Conference on Deep Foundations—Volume 2 23


CPTU F2
Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
Cone Stress, qt (MPa)
0 25 50 75 100 0 200 400 600 800 1,000 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10

DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 28. Results of CPTU F2.

CPTU F3
Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
Cone Stress, qt (MPa)
0 25 50 75 100 0 200 400 600 800 1,000 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10
DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 29. Results of CPTU F3.

24 3rd Bolivian International Conference on Deep Foundations—Volume 2


CPTU G1
Cone Stress, qt (MPa) Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
0 25 50 75 100 0 100 200 300 400 0 1 2 3 4 5
0 5 10 15
0 0 0
0
?

5 5 5 5

10 10 10 10

DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 30. Results of CPTU G1.

9.3 Dilatometer Tests


Figure 31 is a compilation of all DMT results. For graphic and text files with the results of the
dilatometer tests, see the conference web site.

Fig. 31. Compilation of all DMT results.

3rd Bolivian International Conference on Deep Foundations—Volume 2 25


Fig. 32. DMT results, A3.

26 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 33. DMT results, B2

Fig. 34. DMT results, C1.

3rd Bolivian International Conference on Deep Foundations—Volume 2 27


Fig. 35. DMT results, F1.

Fig. 36. DMT results, F1.

28 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 37. DMT results, D1.

Fig. 38. DMT results, E2.

3rd Bolivian International Conference on Deep Foundations—Volume 2 29


Fig. 39. DMT results, G1.

Fig. 40. DMT results, G1.

30 3rd Bolivian International Conference on Deep Foundations—Volume 2


TEST EVALUATION REPORTS

Boreholes and SPT-T

Decourt, L.(1)
(1)
Luciano Decourt Egenharios Ltda., Sao Paulo, Brazil <decourt@decourt.com.br>

Nine SPTs have been carried out at the site of B.E.S.T. Most of them also included torque
measurements and some have been instrumented to provide Enthru energy measurements.
As is well known, for common soils, the SPT provides information usually considered reliable
for foundation design purposes. However, for structured soils, such as soft clays, and saprolitic
and lateritic soils, SPT results are not reliable. The introduction of torque measurements to these
tests, the SPT-T (Décourt and Quaresma Filho 1991; 1994) and the concept of equivalent N, Neq
(Décourt, 1991; 2002) allowed a much better understanding of the behavior of soft clays and
saprolitic soils. However, for lateritic soils, no test, up till know, allows correct evaluation of the
response from the SPT records.
Average N-indices SPTs, for each depth are shown in Figure 1 and in Table 1.

TABLE 1. Average N-index

Fig. 1. Distribution of SPT NSPT-indices and average values.

3rd Bolivian International Conference on Deep Foundations—Volume 2 31


The energy effectively transferred to the roads, called Enthru Energy, was measured. On
average, this energy corresponds to an efficiency of about 44%.
In order to apply methods developed elsewhere to the Santa Cruz de la Sierra soils, it is
absolutely fundamental, that their efficiencies, be well known. For comparison, according to
Décourt (1991) for the soils of the Tertiary Sedimentary Basis of São Paulo, TSBSP, the torque, T,
values, measured in kgf.m (10 kNm) are, on average, 1.2 times the Brazilian NSPT values, here
designated N72, that presents an average efficiency of 72%. The results of torque measurements,
carried out, in the B.E.S.T. site may be also used to assess the efficiency of these tests.
Figures 2 and 3 show correlations between Neq and T values (kNm)/12 x NSPT (field
measured SPT values) for the SPT results of all clayey soils of the B.E.S.T. site (records from the
sandy soils were discarded because their interpretation might be problematic).

EBEST = 0.672 X 72
∵ EBEST ≈ 48.4%

Fig. 2. Correlation between Neq and NSPT. Fig. 3. Correlation between Neq and
NSPT, with regression line through the
origin.

Some tests carried out at this site had energy measurements. Figures 4 and 5 show correlations
between N60 and NSPT.
N60 ≈ 0.769 NSPT.. EBEST = 60 x 0.769 = 46.14 %. Another option is to compare quc (CPT)
values with NSPT. Considering depths down to 17.85 m, the ratio of the average values of qc and
NSPT is:
qc/NSPT = 4.375/10.400 = 0.42 (1)

For the Brazilian SPT, Ei = 72 %, Velloso and Lopes (1996) proposed:

qc/N72 ≈ 0.6 EBEST = 0.42/ 0.6 x 72 = 50.4 % (2)

32 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 4. Correlation between N60 and NSPT. Fig. 5. Correlation between N60 and NSPT,
with regression line through the origin.

Besides, as mentioned above, the sampler had a provision for the use of liners. However it
was not used. Were it to be used, it could yield higher efficiencies than now measured. On basis
of all this reasoning, it appears reasonable to assume for the SPTs carried out at B.E.S.T., an
apparent average efficiency of 48%. So, in order to convert the measured NSPT values to N60, the
field values should be divided by a factor of 1.25. Another important consideration regards fine
sands below the water table. the influence of pore pressure generation also effects NSPT values.
According to Terzaghi and Peck (1948; 1996), for saturated, very dense, fine sands, the measured
N values should be reduced, as follows:

Ncorr = 15 + 0.5(NSPT - 15) (3)

The efficiency of the SPTs considered in Terzaghi and Peck analysis was not known.
However, most probably, they were in the range of 45% - 50%, approximately the same range of
values determined in B.E.S.T. tests.
For dense sands, NSPT ≥ 15, the void ratios were assumed to be lower than those corresponding
to the Critical State, and the tendency was to dilate and to the generate negative pore pressures. As
a consequence, the measured NSPT values might be unrealistically high. For NSPT values lower
than 15, the opposite occurs.
Figure 6 shows the ratio of corrected values of NSPT, divided by NSPT and as a function of NSPT.
The correction proposed by these authors had been extended for NSPT values lower than 15, what
has not been the intention of these authors.
Figure 7 shows the Terzaghi and Peck corrections adapted to the Brazilian SPT (Ei = 72%),
and another correction based on equivalent N values (Neq) are shown (Décourt 1998).
For very loose fine sands, below the water level, the measured NSPT values should be increased
as indicated in these figures.
In some cases, as mentioned by Décourt (1986), unrealistically low NSPT values were the most
likeable explanation for the fact that the capacities of displacement piles, computed using Décourt
and Quaresma method (1978) have been much lower than those provided by loading tests. Once

3rd Bolivian International Conference on Deep Foundations—Volume 2 33


the low NSPT values were corrected as suggested in Figures 6 and 7, the differences between -
calculated and measured values of capacity became negligible.

Fig. 6. Corrections to measured Fig. 7. Variation of Neq and Ncorr


NSPT values (after Decourt (1998). with N72 in fine sand
below the water table.

REFERENCES

Décourt, L., 1986. One case where the SPT failed in the prediction of bearing capacity of piles (in
Portuguese). VII Cobramseg, Porto Alegre, Brazil.
Décourt, L., 1991. Bearing capacity of displacement piles in residual soils on basis of SPT, SEFE
II, São Paulo, Brazil.
Décourt, L. and Quaresma Filho, A.R. 1991. The SPT-CF, an improved SPT, SEFE II, São Paulo,
Brazil.
Décourt, L. and Quaresma Filho, A.R. 1994. Practical applications of the standard penetration test
complemented by torque measurements, SPT-T; Present state and future trends. XIII
ICSMFE, New Delhi, India / XIII CIMSTF, 1994, New Delhi, India.
Décourt, L., 1998. A more rational utilization of some old in situ tests. 1st First International
Conference on Site Characterization – ISC. Atlanta, U.S.A.
Décourt, L., 2002. The concept of Equivalent N, Neq, in engineering practice; Still a postulate or
already a proven reality?( in Portuguese). VII Cobramseg. São Paulo, Brazil.
Terzaghi, K. and Peck, R.B., 1948; 1996. Soil mechanics in engineering practice. John Wiley and
Sons, New York, U.S.A.
Velloso, D.A. and Lopes, F.R., 1996. Foundations (in Portuguese). Coppe-UFRJ Rio de Janeiro,
Brazil.

34 3rd Bolivian International Conference on Deep Foundations—Volume 2


Cone Penetrometer Soundings, SCPTU Results

Robertson, P.K.(1)
(1)
Gregg Drilling and Testing Inc., Signal Hill, California, USA
<PRobertson@GreggDrilling.com>

Figure 1 shows an overlay plot of the SCPTu for each pile location (A1 to G1). In general, the
SCPTu results indicate a complex sequence of interbedded sands and clays. The soils are
reasonable consistent across the site with some variability in layer thickness and consistency.
Figure 2 shows an overlay plot of estimated soil parameters, based on Robertson (2009). The
overlay plots of estimated soil parameters illustrate the variability within each main stratigraphic
unit. Several dissipation tests were carried out at various depths and the time for 50% dissipation
(t50) in the fine-grained soils between depths of about 12 to 16 m varied from 2.5 to 10 minutes
with an average of 4.8 minutes, that is consistent with a silty clay/clayey silt type of soil.

Fig. 1. Overlay plots of SCPTu profiles (A1, B1, C1, D1, E1, F1, and G1) in terms of cone
resistance, qt, sleeve resistance, fs, penetration pore pressures, u2 and SBT index, Ic.

The measured shear wave velocity values (Vs) are generally higher than estimated using the
CPT data based on empirical correlations for young, uncemented soils. This would suggest that
the soils have significant microstructure due to either aging and/or cementation effects (Robertson,
2016). Figure 3 includes the measured values of Vs (shown in red) for location G1, as an example.
Other SCPTu locations show a similar trend of high Vs values. Robertson (2016)

3rd Bolivian International Conference on Deep Foundations—Volume 2 35


Fig. 2. Overlay plots of SCPTu profiles (A1, B1, C1, D1, E1, F1, and G1)
in terms of estimated soil parameters.

36 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 3. SCPTu G1 comparing measured to estimated Vs profile.

Fig. 4. SCPTu G1 normalized SCPTu data and normalized rigidity index (K*G),
based on Robertson (2016).

3rd Bolivian International Conference on Deep Foundations—Volume 2 37


had suggested that the SCPTu data can be used to calculate a normalized rigidity index, K*G, based
on the ratio of Go/qn. When K*G exceeds 330, the soils have significant microstructure and can be
expected to be stiffer at small strains compared to young, uncemented soils at a similar in-situ
state. Figure 4 shows that much of the profile has K*G > 330.
Robertson (2016) suggested that when soils have significant microstructure (i.e. K*G > 330)
traditional empirical correlations may not be reliable. Based on traditional empirical correlations,
the fine-grained soils appear to be overconsolidated, where the estimate OCR is variable but
generally around 5 or more over much of the depth investigated. This apparent OCR maybe due to
microstructure and not directly a result of any stress history. The high positive excess pore
pressures measured during the SCPTu suggest a more contractive soil behavior at large strains
which is inconsistent with the relatively high apparent OCR values. This contradiction in estimated
soil behavior appears to be due to light cementation that results in high small strain stiffness
(reflected in Vs) compared to contractive positive pore pressures during cone penetration. The high
measured Vs-values indicate that the soils are likely stiffer at smaller strains than would normally
be expected, that could result in slightly stiffer load-settlement response for some piles, depending
on the level of disturbance induced during pile installation.

REFERENCES
Robertson, P.K., 2009. Interpretation of Cone Penetration Tests – a unified approach. Canadian
Geotechnical Journal. 27(1) 151-158, 10.1139/t90-014
Robertson, P.K., 2016. Cone Penetration Test (CPT)-based soil behavior type (SBT) classification
system–an update. Canadian Geotechnical Journal, 53(12)1910-1927, 10.1139/cgj-2016-0044

38 3rd Bolivian International Conference on Deep Foundations—Volume 2


Dilatometer Tests

Marchetti, D.(1)
(1)
Studio Prof. Marchetti s.r.l., Rome, Italy <diego@marchetti-dmt.it>

The SDMT results indicate an alternation of sands and clays, with a reasonable homogeneity along
all the test locations. In particular, the material index ID indicates two main layers of sand (5-12 m
and below 17 m) and two main layers of finer material (2-5 m and 12-17 m).
The top clayey layer is characterised by KD ≈ 2-3, indicating a normally consolidated or
slightly overconsolidated state.
The deeper clayey layer exhibits higher values of KD, with OCR with values of about 3-5 in
the northern and southern locations and even higher in the central location C1 and F1 (OCR up to
7-8). The undrained shear strength Cu also shows an increase in the same central locations, with
values up to about 250 kPa. The higher mechanical characteristics in the central locations are
confirmed by the higher values of Vs and G0, obtained with independent measurements from the
seismic instrumentation.
The sandy layer between 5-12 m is characterized by a high variability of KD and MDMT along
the various test locations. Similar variability is observable in the qt profiles of the CPT. The
average ratio of MDMT/qt is between 5-8, suggesting an approximate estimation of OCR ≈ 1
(Monaco et al. 2014).
The P2 values in the sandy layers confirm the depth levels of the water table in the locations
where the C readings were performed.
Figure 1 shows that the measured Vs values are significantly higher than their estimation from
the DMT parameters (Amoroso 2014), suggesting that the soil may be affected by aging,
cementation, or structure.

Fig. 1. Measured values vs. depth.

3rd Bolivian International Conference on Deep Foundations—Volume 2 39


40 3rd Bolivian International Conference on Deep Foundations—Volume 2
Pressuremeter Tests

Frank, R.(1)
(1)
Ecole Nationale des Ponts et Chaussées, Champs-sur-Marne, France.
<frank@roger.frank.enpc.fr>

TEAM pressuremeter expansion tests (PMTs) were performed in 5 boreholes on the B.E.S.T. site,
located at A3, B2, C1, D1, and E2. In the 5 boreholes, the tests were performed at the same 3
depths: 5.54 m, 9.04 m (with the exception of C1, where the depth was 8.94 m), and 12.04 m.
Altogether 15 expansion tests are thus available. The TEXAM tests were performed with the
“equal-volume increment method” (also called “volume-controlled” tests), following ASTM D
4719-07 standard (Procedure B). The TEXAM probe is made of one cell with length/diameter ratio
L/D = 6.2 (NX probe, with D = 73.8 mm and L = 0.46 m), ensuring negligible end effects. The
initial volume of the probe is 1,968 cm3.
Note that ‘conventional’ Ménard type prebored PMTs are usually carried out with a probe
encompassing three cells and using the “equal-pressure increment method” (Procedure A in ASTM
D 4719-07; see also EN ISO standard, AFNOR, 2015). According to Briaud (1992), the results
obtained from TEXAM tests and Ménard tests are quite similar.
For the 5 boreholes, mud rotary drilling with a drag bit was used. The diameter of the drag bit
was around 74 mm, very near the diameter of the probe (Viscarra 2017). According to the site
records, there was no need to force the probe into the ground. Thus, it seems that the size of the
preborehole was optimal (only slightly larger than the size of the probe), ensuring good quality
expansion tests. Indeed, the observation of the expansion curves (plotted in Section 9.4 as pressure,
p, vs. radial strain, � r/r0) confirm the good quality of the tests.
Apart from the expansion curves, obtained from the raw injected volume and raw pressure
readings corrected for volume loss, pressure loss, and hydrostatic pressure (see ASTM standard or
EN ISO standard), the other main results from the interpretation of the PMTs are:

- the pressuremeter modulus, EM ;


- the creep pressure, pf ;
- and the conventional limit pressure, pL.

Figure 1 is an example of the interpretation of a Ménard expansion test following the EN ISO
standard. The expansion curve is plotted as injected volume, V, vs. pressure, p. The modulus, EM,
is determined from the slope between the points (p1, v1) and (p2, v2). The limit pressure, pL, is
determined by extrapolation of the expansion curve to the doubling of volume of the cavity, that
is, for an injected volume equal to Vs + 2V1 (where Vs is the initial volume of the pressuremeter
probe and V1 is the volume injected in order to catch back the initial wall of the cavity). Several
extrapolation methods are used, such as the “inverse volume method” and the “double hyperbola
method”. The creep pressure, pf, is determined from the creep curve (difference in injected
volumes recorded at 60 s and 30 s for each pressure increment � V60/30 = V60 –V30).

3rd Bolivian International Conference on Deep Foundations—Volume 2 41


1200

VS + 2V1
1000

800

EM = 23 MPa
600

400

Pl = 1,7 MPa
200 p2
pf
p1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8

Fig. 1. Interpretation of a Ménard expansion test following the EN ISO standard (document:
APAGEO).

For the present interpretation of the PMTs performed on the B.E.S.T site, the proposed
pressuremeter modulus EM is an upper bound, as the highest slope of the expansion tests is chosen
(Viscarra 2017) – see the example on Figure 2. The limit pressure is obtained by extrapolation of
the 1/V curve to the doubling of the volume of the cavity, that is, for V > 2000 cm3 or � r /r0 around
40% (“inverse-volume method”). It is to be noted that, in the present tests, the maximum injected
volumes are rather low: they vary between 431 cm3 (� r/r0 = 10.4%) and 917 cm3 (� r/r0 = 21%).
More often they are around 600 cm3 to 700 cm3. The extrapolation would certainly be more reliable
if the maximum injected volumes would be about 1,000 cm3 or more. In the example shown on
Figure 3, the maximum injected volume is 880 cm3.

Fig. 2. Expansion curve of Test No. 2 at location C1 of B.E.S.T. site – see Section 9.4.

42 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 3. Determination of the limit pressure pL by the “inverse volume method”.
Test No. 2 at location C1 on B.E.S.T. site, PMT C1– see Section 9.4.

Another interpretation has been carried out by Reiffsteck (2017), using the ISO EN standard
for Ménard PMTs (AFNOR 2015) for deriving the corrected expansion curves, as well as the
‘user’s choice’ for deriving EM and the double hyperbola method for deriving pL – see example in
Figure 4. This interpretation gives a picture which is overall consistent with the results given in
Section 9.4, but it leads to moduli EM usually lower and limit pressures pL usually larger than those
presented in Section 9.4.
According to the borings on the B.E.S.T. site, it is clear that the tests at 9 m depth are all in
the sand layer, while the tests at 5.5 m depth and 12 m depth are either in a predominantly sand
layer or a silty sand and sandy silt layer, at the interface of these two layers, or at the interface of
one of these two layers with a clay layer. For these depths, it is not possible to be entirely sure
about the nature of the layers.
In the sand layer at 9 m depth, the limit pressures, pL, range between 670 kPa and 990 kPa,
and the values of EM/pL are between 7 and 11. These values are quite consistent and are typical of
medium dense sands. For the silty sand and sandy silt apparently found at A3, C1 and D1, at 5.5 m
depth, the limit pressures pL are between 140 kPa and 400 kPa. There are quite low values, usually
measured in very loose to loose sands or silts. EM/pL are between 7.7 and 8.1, matching more

3rd Bolivian International Conference on Deep Foundations—Volume 2 43


values for stiffer soils. With regard to clay, only the test at location E2 and 5.5 m depth might be
relevant, but it is located very near a sand and silty sand layer, if not at the interface.

Fig. 4. Interpretation of test No. 2 at location B2 on B.E.S.T. site, following EN ISO standard
(Reiffsteck 2017).

The values pL = 240 kPa and EM/pL = 8.6 would then mean that the clay is very soft. The pL
value might also show loose sand, but then again the value EM/pL is too high.
An unload-reload loop has also been incorporated into each expansion test on the B.E.S.T.
site. The corresponding moduli, Eunload and Ereload are several times larger than the ‘first loading’
modulus, EM, as also expected. Eunload and Ereload can sometimes be used as a measure of small
deformation moduli or a sort of Young’s modulus (to be used, for example, in formulae, tables,
and charts of elastic solutions). This approach is a rather delicate matter, because the values
obtained depend on a number of factors difficult to quantify.
For the prediction of the behaviour of the piles tested on the site, the PMT results can be very
useful. Indeed, the bearing capacity can be derived from the limit pressures pL and the load-
settlement curve can be derived from the pressuremeter moduli EM, by applying the so-called
‘Frank and Zhao’ t-z curves (see Burlon et al. 2014 and Abchir et al. 2016). Unfortunately, the
PMT tests were only performed at 3 depths. Nevertheless, by using correlations between the PMT
results and the results of other in-situ tests performed on the site, it may prove possible to build up
a full pressuremeter ‘picture’ allowing to predict the behaviour of the piles on the site from the
pressuremeter parameters EM and pL.

44 3rd Bolivian International Conference on Deep Foundations—Volume 2


REFERENCES

Abchir Z., Burlon S., Frank R., Habert J., and Legrand S., 2016. t-z curves for piles from
pressuremeter test results. Géotechnique 66(2) 137-148 [doi:10.1680/jgeot.15.P.097]
AFNOR, 2015. Geotechnical investigation and testing - Field testing - Part 4: Ménard
pressuremeter test. European and International Standard, NF EN ISO 22476-4, May 2015,
51 p.
ASTM D4719-07, Standard Test Methods for Prebored Pressuremeter Testing in Soils (Withdrawn
2016), 9 pages, ASTM International, West Conshohocken, PA, 2007 [doi:10.1520/D4719-
07]
Briaud J-L., 1992. The pressuremeter. Balkema, Rotterdam, 322 p.
Burlon S., Frank R., Baguelin F., Habert J., Legrand S., 2014 Model factor for the bearing capacity
of piles from pressuremeter test results — Eurocode 7 approach, Géotechnique 64(7) 513–
525 [doi:10.1680/geot.13.P.061].
Reiffsteck P., 2017. Interpretation of PMTs performed at Santa Cruz de la Sierra, Bolivia. Private
communication.
Viscarra F., 2017. Private communication by e-mail, 19 January 2017.

3rd Bolivian International Conference on Deep Foundations—Volume 2 45


46 3rd Bolivian International Conference on Deep Foundations—Volume 2
Interpretation of Seismic Tests

Massarsch, K.R.(1), Noel Perez(2), Terceros A. M.(2) and Terceros H. M. A.(2)

Geo Risk & Vibration AB, Bromma, Sweden. <rainer.massarsch@georisk.se>


(1)

(2)
Incotec S.A., Santa Cruz de la Sierra, Bolivia <mta@incotec.cc>, <math@incotec.cc>
<noel.perez@incotec.cc>

ABSTRACT. Comprehensive geotechnical and seismic investigations have been carried out at
the B.E.S.T. site. The following tests were performed: seismic refraction, surface wave
measurements (MASW) and seismic down-hole tests (one and two sensors). The investigations
offer a unique opportunity to compare the results of the different seismic tests at a relatively
homogeneous site. The effect of strain softening on the shear modulus is accounted for by the
introduction of a modulus reduction factor. Geotechnical parameters such as plasticity index, void
ratio and degree of saturation influence the modulus reduction factor. A concept is presented which
makes it possible to estimate the soil modulus at large strain. The results of shear wave velocity
measurements by different methods are compared. The shear modulus and elastic modulus at large
strain (static modulus) is estimated.

1. GEOTECHNICAL SETTING
The geology of the B.E.S.T. site is characterized by a sedimentary basin. The soil deposit is the
result of a sedimentation-erosion-sedimentation process, dominated by fine to medium sands with
intermittent layers of silt, clay or clayey sand. The results of SPT investigations and laboratory soil
classification at test point C1 are shown in Figure 1.
SPT N-INDICES
H-C1 WATER CONTENT (%) SOIL TYPE FRACTIONS (%) (blows/0.3 m)
0 5 10 15 20 25 30 35 40 45 50 0 20 40 60 80 100 0 10 20 30
0 0 0
GW
2 2 2
SAND
PL wn LL CLAY and SILT
4 4 4

6 6 Loose Gravel 6

8 8 8
SAND
DEPTH (m)

10 10 10

12 12 Clay 12
Silty Sand
14 14 14
CLAY
SAND
16 16 and 16
SILT
18 18 18

20 20 20

22 22 22

Fig. 1. Results of SPT investigation and soil classification in borehole C1.

3rd Bolivian International Conference on Deep Foundations—Volume 2 47


The upper about 20 m thick soils consist of normally consolidated layers of clay, silt, and
sand, in various combination and thickness. The upper about 5 to 6 m consists of loose silt and
sand. Hereunder lies a 6 to 7 m layer of compact silt and sand. At about 11 m depth exists an about
1 m thick layer of soft silty clay followed by an about 1 m thick layer of compact sand. Below
about 12 m depth, the profile alternates between about 2 m thick layers of compact to dense silty
sand and about 2 m thick layers of loose sand. The groundwater table at the site ranges seasonally
between the ground surface and about 0.5 m depth.
Sleeve Friction, fs (kPa) Pore Pressure (kPa) Friction Ratio, fR (%)
Cone Stress, qt (MPa)
0 25 50 75 100 0 200 400 600 800 1,000 0 1 2 3 4 5
0 5 10 15
0 0 0
0

5 5 5 5

10 10 10 10

DEPTH (m)
DEPTH (m)
DEPTH (m)
DEPTH (m)

15 15 15 15

20 20 20 20

25 25 25 25

Fig. 2. Results of CPTU at tests pile location C1.

The results of the borehole records and the CPTU from Location C-1 show reasonable
agreement. However, the CPTU provides significantly more detailed information. The cone stress
shows four general soil layer formations (0–6 m; 6–12.5 m, 12.5-16.5 m and 16.5-22 m), and the
pore water pressure measurements add additional valuable information. Down to about 12.5 m,
the soil is relatively free-draining, with occasional fine-grained layers. Below 15.5 m, the deposit
changes to mainly fine-grained soil. The relatively stiff/dense soil layer between 12.5 and 16.5 m
is characterized by negative pore water pressure, indicating a dilative behavior. Different types of
seismic measurements were performed at individual locations (SCPT and SDMT) as well as along
three profiles. Figure 3 shows and overview of the B.E.S.T. site with test location and seismic
profiles (MASW and refraction).

Fig. 3. Test area showing of test piles (Ax-Hx) and seismic profiles (A, B and C).

48 3rd Bolivian International Conference on Deep Foundations—Volume 2


Seismic tests (MASW and refraction) were performed along three approximately 80 m long
profiles (A, B and C). At each test pile location (Ax through Hx), geotechnical (SPT, CPTU, DMT)
and seismic tests (SCPT and SDMT) were performed at a distance of 0.8 m from the center of the
pile, as illustrated at Location A-1 shown in Figure 4.

Fig. 4. Location of geotechnical (SPT, DPSH, CPTU and DMT) and seismic tests (SCPT and
SDMT), illustrated at Location A-1.

2. SEISMIC TESTING METHODS


Seismic methods have been used extensively within earthquake engineering but are also used
increasingly for the solution of geotechnical problems. An important reason has been the rapid
development of powerful electronic measurement systems suitable for use under difficult site
conditions. Further, the analytical capacity of programs, which can be operated on conventional
computers has increased significantly. In the following, the four seismic methods used as part of
the investigation are briefly described. For more details, reference is made to publications by e.g.,
Richart et al. (1970), Santamarina et al. (2001), and Stokoe et al. (2004).
There are two types of seismic body waves, pressure or compression waves (P-waves) as well
as shear waves (S waves), and seismic sensors react to both. The P-wave always arrives first. In
soils below the groundwater table, the P wave typically travels 2 or more times faster than the S-
wave, so separation of the two body waves is easy. Above the water table, however, the difference
is small and separation of P- and S-waves may be very difficult, requiring specialized techniques.
However, the most significant difference between P- and S-waves is that S-waves are reversible.
Therefore, using a source that can produce shear waves of opposite polarity facilitates the
identification of S-waves. Since shear waves travel through the skeletal structure of the soil at very
small strains, one can apply simple elastic theory to calculate the average elastic small strain shear
modulus, over the length interval of measurement, as the mass density times the square of the shear
wave velocity. Thus, the shear wave velocity relates directly to stiffness (Massarsch 2004) and
may also be used to estimate liquefaction susceptibility in young uncemented sands (Youd et al.
2001).

3rd Bolivian International Conference on Deep Foundations—Volume 2 49


2.1 Seismic Refraction Method
Seismic refraction has been the most commonly used method in geophysical site exploration and
earthquake engineering. However, due to its limitations, it has been replaced by more suitable
methods for the solution of geotechnical and soil dynamics problems. In this method, a series of
receivers, usually geophones, are placed in a linear array. An energy source (hammer blow or small
explosive charge) is then used to generate compression waves, cf. Figure 3. Vertically sensitive
geophones are installed at increasing distance, d, from the source, P to record the arrival of
compression waves which propagate along soil or rock layers of varying thickness H, cf. Figure 3.
The first arrival of the waves is recorded by the geophones. A fundamental assumption of the
refraction method is that the compression wave velocity increases with depth. This requirement
imposes an important limitation of the method, as in many geological formations, soft or loose soil
layers with lower velocity can occur below a high-velocity layer.
At geophones located close to the point of impact, such as point G1, the direct waves reach first.
At the points located away from the source, the refracted waves arrive earlier than the direct waves,
Figure 5.

Fig. 5. Arrangement of vertically sensitive sensors (G) and source (P) at seismic refraction test.

From the recorded first arrival of compression waves, the time-distance relation can be
plotted, Figure 6. The time, t, of arrival of the first impulse at various geophones is taken as
ordinate and the distance, d, of the geophones from the source P is taken as abscissa. Velocity in
any layer is equal to the reciprocal of the slope of the corresponding line. The slopes of the various
lines are determined, from which the corresponding velocities are computed.

50 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 6. Time-distance plot of seismic refraction test, cf. Figure 3.

After the determination of velocities at different layers, its depth can be calculated from the
following equations:

d1 V −V
H1 = �V2 +V1 (1)
2 2 1

d2 V −V
H2 = 0.85 H1 + �V3 +V2 (2)
2 3 2

It should be noted, that the above case is limited to level ground and gradually increasing wave
velocities. More sophisticated, computer-based, analytical models are available.

2.2 Seismic Down-hole Method


Seismic cross-hole and down-hole methods have been widely used soil dynamics and geotechnical
earthquake engineering since the beginning of the 1960s. The down-hole method is most widely
used today due to the incorporation of seismic sensors in the cone penetration test—the seismic
cone penetrometer (SCPT). The seismic down-hole method has been described in the geotechnical
literature, (Robertson et al. 1986). Standards and guidance documents have been developed, such
as the ISSMGE Guideline “Seismic cone downhole procedure to measure shear wave velocity”,
which is attached to this document, (Butcher et al. 2015).
During a pause in cone penetration, a shear wave can be created at the ground surface that will
propagate into the ground on a hemispherical front and a measurement made of the time taken for
the seismic wave to propagate to the seismometer in the cone. By repeating this measurement at
another depth, one can determine, from the signal traces, the interval time and so calculate the
average shear wave velocity over the depth interval between the seismometers. A repetition of this
procedure with cone advancement yields a vertical profile of vertically propagating shear wave
velocity. The general arrangement of SCPT equipment is shown in Figure 7.

3rd Bolivian International Conference on Deep Foundations—Volume 2 51


a) SCPT with single receiver b) SCPT with dual receivers

Fig. 7. Schematic diagram of SCPT with single and dual seismic sensors, (Butcher et al. 2015).

Figure 8 illustrates the execution of a seismic down-hole test (SDMT) at the B.E.S.T. site. A
horizontally polarized shear wave is generated by striking a steel plate (loaded by the CPT rig). A
trigger activates the recording equipment that then displays the time based signal trace received by
the seismometer.

Fig. 8. Execution of a seismic down-hole test (SDMT) at the B.E.S.T. site, courtesy Incotec.

52 3rd Bolivian International Conference on Deep Foundations—Volume 2


An example of a pair of signals is shown in Figure 9. With reversed image traces, the first
major cross-over can be taken as the “reference” arrival, or one trace may be used and an arrival
pick made visually by an experienced operator. Alternately, a cross-correlation procedure may be
used to find the interval travel time using the wave traces from strikes on the same side at
successive depths. The latter technique is more complex, but eliminates the arbitrary visual pick
of arrival time and is necessary if symmetry of reverse wave traces is lacking.

Fig. 9. An example of oppositely polarized shear wave traces with clear crossover of traces
showing the interval time T2 – T1. (Butcher et al. 2015).

The seismic dilatometer SDMT is the combination of the flat dilatometer with an add-on seismic
module for the measurement of the shear wave velocity VS. The measurement system is similar to
that of the SCPT. However, the SDMT uses two seismic receivers and the true-interval time can
be measured, which enhances the repeatability of the VS measurements. The seismograms recorded
by the two receivers, amplified and digitized at depth, are transmitted to a PC at the surface that
automatically calculates the delay using the cross-correlation algorithm, Figure 10.

Fig. 10. Example of seismic record and re-phased signal using cross-correlation (from DMT
pamphlet).

3rd Bolivian International Conference on Deep Foundations—Volume 2 53


2.3 Surface Wave Methods
The most common approach used in geotechnical and earthquake engineering is called the
spectral-analysis-of-surface-waves (SASW) method. The method was developed at the University
of Texas at Austin and reported by Stokoe et al. (1989). Rayleigh wave energy is generated at one
point and the resulting vertical surface motion is recorded at various distances (receiver points)
away from the source, similar to the test set-up shown in Figure 3. Measurements are performed
at multiple source-receiver spacings along a linear array. The phase shift versus frequency
relationship is measured for surface waves propagating between the receivers for each receiver
spacing. The result is a plot of phase velocity versus frequency for a given receiver spacing, called
an individual dispersion curve. An iterative forward modeling procedure or an inversion analysis
algorithm is used to determine a shear-wave velocity profile by matching the field dispersion curve
with the theoretically determined dispersion curve.
The SASW method uses the apparent phase velocity dispersion curve along with source and
receiver locations in the forward modeling or inversion analysis. The dynamic stiffness matrix
method, which is the forward modeling algorithm used in the matching or inversion process, can
simulate the apparent phase velocity specific to the source receiver configuration. The inversion
analysis based on apparent phase velocities and the dynamic stiffness matrix method are key
features of the SASW method, which improves the reliability and accuracy of the shear-wave
velocity profile.
In the MASW method (Park et al., 1999), a large array of time traces is measured using a
swept-sine vibratory source or an impulsive hammer. The basic field configuration and acquisition
procedure for the MASW measurements is generally the same as the one used in conventional
common midpoint (CMP) body-wave reflection surveys. In the MASW method, the dispersion
curve can be determined in two approaches: the swept-frequency record approach and the
frequency-wave number spectrum approach.
The MASW method uses only the fundamental mode for the inversion analysis. For the site
with abnormal dispersive dispersion curve, in which phase velocities increase with increasing
wavelength, the fundamental mode alone may be enough to resolve the layer stiffness reliably. To
make the MASW method a reliable exploration method, it is crucial to incorporate higher modes
as well as the fundamental mode in the inversion analysis. Recently, an effort to use higher modes
in the inversion analysis was made by Kansas Geological Survey (Pak et al. 1999).
Note the important advantage of surface wave measurements (SASW and MASW) that reliable
measurements can be performed even when wave velocities decrease with depth, or soft layers are
overlain by stiff layers.

3. INTERPRETATION OF SEISMIC TESTS


3.1 Small-strain Shear Modulus
The small-strain shear modulus, Gmax can be determined from the following relationship

𝐺𝐺𝑚𝑚𝑚𝑚𝑚𝑚 = 𝑉𝑉𝑆𝑆2 𝜌𝜌 (3)

where ρ is the bulk density of the soil. Gmax is determined at very low shear strain, typically lower
than 10-5 (10-3 %). At such a low strain level, excess pore pressure is not generated and Gmax reflects
fundamental soil behavior. As has been pointed out by Massarsch (2004), during seismic tests at
shear strain level < 10-3 %, the rate of loading (straining rate) is surprisingly slow and comparable
to that of static laboratory tests. This important aspect has been confirmed by a comparison of

54 3rd Bolivian International Conference on Deep Foundations—Volume 2


resonant column tests, performed at vibration frequencies of 30 to 35 Hz, and static torsional shear
tests, Drnevich and Massarsch (1979). For practical purposes, the effect of strain rate on medium
dense and dense granular soils can be neglected up to a strain level of approximately 0.1%.
The results of a resonant column on an undisturbed, reconsolidated sample of clayey sand is
shown in Figure 11, Drnevich & Massarsch (1979). Below 10-3 % shear strain the shear modulus
appears to be unaffected by shear strain (and thus strain rate). However, when shear strains exceed
10-3%, the shear modulus decreases. At 10-1% shear strain, the shear modulus of the clayey sand
is only about 30% of the maximum value.

Fig. 11. Variation of shear modulus with shear strain determined from torsional resonant column
test, after Drnevich and Massarsch (1979).

Based on extensive resonant column tests, Hardin (1978) suggested that the small-strain shear
modulus, Gmax of sand can be estimated from the following relationship

(4)

where: e = void ratio, σ’m = mean effective stress and σr = reference stress (100 kPa). The mean
effective stress σ’m is defined as

(5)

where: σ’v = vertical effective stress, K0 = coefficient of lateral earth pressure at rest. Even if the
horizontal stress (and thus K0) are not known, it is preferable to estimate the coefficient of
horizontal earth pressure at rest, K0 based on engineering judgment than to neglect the significance
of horizontal effective stress. Hardin (1978) found that, for granular soils, the overconsolidation
ratio, OCR has little or no influence on Gmax. In Figure 12, the variation of the small-strain shear
modulus, Gmax is shown for different values of the void ratio, e as a function of the mean effect

3rd Bolivian International Conference on Deep Foundations—Volume 2 55


stress, σ’m. It is assumed that the groundwater table is at the ground surface, the coefficient of
lateral earth pressure at rest, K0, is 0.5, and the bulk density, ρ, is 2000 kg/m3.

Fig. 12. Variation of the small-strain shear modulus with mean effective stress for different
values of void ratio, cf. Equation (5). The ground water level is assumed at the ground surface.

3.2 Modulus Degradation of Fine-grained Soils


The shear modulus decreases with increasing shear strain level, cf. Figure 11. The static shear
modulus, Gs (defined as shear modulus at a shear strain level of approximately 0.5 % shear strain,
which corresponds to working load at a factor of safety, FS > 1.5) can be estimated from the
following relationship

𝐺𝐺𝑠𝑠 = 𝑅𝑅𝑀𝑀 𝐺𝐺𝑚𝑚𝑚𝑚𝑚𝑚 (6)

where: RM = modulus reduction factor, Gmax = shear modulus at small strain (<10-3%). The
modulus reduction factor, RM of fine-grained soils has been investigated by Massarsch (2004).
Based on the evaluation of extensive resonant column test data, a relationship was found which
describes the variation of the normalized shear modulus is shown as a function of shear strain, for
different values of PI, Figure 13. It is apparent that shear modulus degradation is more pronounced
with decreasing plasticity index, PI.
The effect of basic soil parameters on the modulus reduction factor, RM of silts and sands,
such as plasticity index, PI, void ratio, e, and degree of saturation, Sr, has been investigated by
Massarsch (2015).
A robust relationship for fine-grained soils between RM at 0.5 % shear strain and plasticity
index, PI, has been proposed by Massarsch (2004). The relationship between RM and PI is shown
in Figure 14 for PI values ranging from 0 to 100 %. The following relationship between the
modulus reduction factor, RM, and plasticity index, PI, is obtained

(7)

56 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 13. Variation of the normalized shear modulus as a function of shear strain for different
values of PI, Massarsch (2004).

Fig. 14. Relationship between modulus reduction factor, RM and plasticity index, PI
including data from Table 2 and results by Massarsch (2015).

For granular soils, the dependence of RM on void ratio, e, is shown in Figure 15, from which
the following relationship between void ratio, e, and the modulus reduction factor, RM, is obtained

𝑅𝑅𝑀𝑀 = 0.111𝑒𝑒 + 0.063 (8)

3rd Bolivian International Conference on Deep Foundations—Volume 2 57


Figure 15. Variation of modulus reduction factor RM (0.5 % shear strain) with void ratio in silt
and sand, (Massarsch 2015).

In spite of some scatter in the data, it is apparent that the modulus reduction effect is more
pronounced in dense (low void ratio) than in loose (high void ratio) granular soils. On average,
Gmax decreases at a shear strain level of 0.5 % to between 10 and 15 % of the maximum value.
For a void ratio between e = 0.3 and 0.8, RM varies between 0.096 and 0.152. On average, in
medium dense (compact) sand with a void ratio e = 0.60, RM is 0.13. The relative density of sands
can be approximately characterized by the ranges of void ratio shown in Table 2.

TABLE 2. Approximate range of values for void ratio in sand with different densities (Massarsch
2015).

Density Void Ratio, e


Very dense 0.35 – 0.45
Dense 0.45 - 0.55
Compact 0.55 – 0.65
Loose 0.65 – 0.75
Very loose 0.75 – 0.85

The dependence of the modulus reduction factor on the degree of saturation is shown in Figure 16.
There is scatter in the data at high degree of saturation, but the trend shows that Sr has only a slight
effect on RM.

58 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 16. Variation of modulus reduction factor RM at 0.5 % shear strain with degree of
saturation.

From Figures 16, the effect of the degree of saturation on the modulus reduction factor can be
determined from the following equations

𝑅𝑅𝑀𝑀 = 0.0003 𝑆𝑆𝑟𝑟 + 0.1069 (9)

A slight increase of the modulus reduction factor appears to occur with increasing degree of
saturation. For most soils the average value of RM can be assumed to be 0.13, a value similar to
that of the void ratio. However, for most practical purposes, the influence of Sr on RM can be
neglected.

3.3 Relationship between Moduli


It is recommended to determine the elastic (Young’s) modulus, E, and the constrained modulus,
M, from the shear modulus G (at strain level ≈ 0.5 % shear strain).

𝐸𝐸 = 2(1 + 𝜈𝜈)𝐺𝐺 (10)

(1−𝜈𝜈) 2(1−𝜈𝜈)
𝑀𝑀 = (1−2𝜈𝜈)(1+𝜈𝜈)E = (1−2𝜈𝜈)
𝐺𝐺 (11)

It should be noted that Poisson’s ratio, ν is strain dependent and increases with increasing strain
level. The ratio between different moduli for a range of values of Poisson’s ratio, ν is shown in
Table 3.
The variation of the elastic modulus, E as a function of mean effective stress can be determined
by substituting Equation (5) into Equation (11). For sand an average value RM = 0.13 is chosen.
Assuming that the ground water table is located at the ground surface, Poisson’s ration, ν = 0.30,
unit weight ρ = 20 kN/m3 and coefficient of lateral earth pressure at rest, K0 = 0.5, the elastic
modulus, E at 0.5% shear strain (i.e. “static modulus”) can be determined. The variation of elastic

3rd Bolivian International Conference on Deep Foundations—Volume 2 59


modulus as a function of mean confining stress and for different values of void ratio is shown in
Figure 17.

TABLE 3. Modulus ratio for values of Poisson’s ratio, ν, cf. Eq. (10) and (11).

Poisson's ratio E/G M/G M/E


0.25 2.50 3.00 1.20
0.30 2.60 3.50 1.35
0.33 2.66 3.94 1.48
0.40 2.80 6.00 2.14
0.49 2.98 51.00 17.11

For example, in granular soils (ν = 0.30), the elastic modulus E = 2.6 G and the confined
modulus M = 3.5 G. The variation of the elastic modulus, E as a function of mean effective stress
can be determined by substituting Equation (5) into Equation (11). For sand an average value RM
= 0.13 is chosen. Assuming that the ground water table is located at the ground surface, Poisson’s
ration, ν = 0.30, unit weight ρ = 20 kN/m3 and coefficient of lateral earth pressure at rest, K0 = 0.5,
the elastic modulus, E at 0.5% shear strain (i.e. “static modulus”) can be determined. The variation
of elastic modulus as a function of mean confining stress and for different values of void ratio is
shown in Figure 17.

Fig. 17. Dependence of elastic modulus, E, at 0.5% shear strain on mean confining stress as a
function of void ratio, e.

4. RESULTS OF SEISMIC MEASUREMENTS


As described above, three different types of seismic measurements have been performed,
comprising: seismic refraction, surface wave measurement (MASW) and seismic downhole tests
(SCPT and SDMT). Seismic refraction and MASW measurements and interpretation of data was
carried out by WARNES, Santa Cruz and SCPT and SDMT tests by Incotec, respectively. A large

60 3rd Bolivian International Conference on Deep Foundations—Volume 2


number of tests have been which will be made available as part of the B.E.S.T. project. Typical
results along the test area have been chosen and are presented below.

4.1 Seismic Refraction


The seismic refraction measurements are two-dimensional along a 92 m long profile, cf. Figure 2.
Compression waves were generated by means of hammer blows on the ground surface. The results
are presented in a tomographic image of compression waves (Vp) and are compared to the cone
stress measurement in test point C1, Figure 18.
Cone Stress, qt (MPa)
0 5 10 15
0

10

DEPTH (m)
15

20

25

Fig. 18. Result of seismic refraction measurement, variation of compression wave velocity, VP
along Line B and comparison with CPT C1, cf. Figure 2.

The image indicates a relatively homogenous medium with generally horizontal layers and a
final refractive boundary at approximately 20.0 m. The first layer down to approximately 5 m
depth shows a P-wave velocity of 400 to 750 m/s. The subsequent layer (5 to 8 m) has a P-wave
velocity of 750 to 1,300 m/s. Below follows a layer (8 to 20 m) with gradually increasing P-wave
velocity (1,300 m/s to 1,900 m/s). At 20 m depth, a stiff boundary was detected.
As mentioned above, seismic refraction cannot detect soil layers, which are overlain by a high
velocity layer. Thus it is not surprising, that the compressible layer between 12 and 16 m is not
detected. The groundwater table is located close to the ground surface. In water-saturated soils, the
P-wave velocity is approximately 1,450 m/s. Thus, the P-wave velocity down to about 10 m depth
is underestimated. It can be concluded that seismic refraction gives a qualitative representation of
soil strata, but is not suitable for geotechnical applications.

4.2 Surface Wave Measurement


Surface wave measurements can be carried out in soil deposit with varying wave velocity profiles.
Multichannel Analysis of Surface Waves (MASW) was carried out along two profiles, A and C,
cf., Figure 3. Rayleigh waves were generated by a hammer blow and surface waves were recorded

3rd Bolivian International Conference on Deep Foundations—Volume 2 61


along two profiles. The results of the MASW investigations are shown in Figure 19 and 20. Also
shown is the variation of cone stress with depth, cf. Figure 3.
It should be noted that the distance between Locations A and C is 6 m. Thus, the two profiles
should give similar results, considering the homogeneity of the site. Both profiles show similar
results, with shear wave velocities increasing generally with depth, from about 150 m/s near the
ground surface, to 180 m/s at 5 m depth. The stiffer layer between 5 and 12 m is also detected,
with an average shear wave velocity of 200 m/s, followed by a layer with higher wave velocity.
However, in profile C, a layer is detected with lower velocity (200 m/s) is found embedded in the
soil layer with higher velocity (220 m/s).
Cone Stress, qt (MPa)
0 5 10 15
0

10

DEPTH (m)
15

20

Fig. 19. Result of MASW measurement, variation of shear wave velocity, VS along Line A and
comparison with CPT C1, cf. Figure 2.
Cone Stress, qt (MPa)
0 5 10 15
0

10
DEPTH (m)

15

20

Fig. 20. Result of MASW measurement, variation of shear wave velocity, VS along Line C and
comparison with CPT C1, cf. Figure 2.

62 3rd Bolivian International Conference on Deep Foundations—Volume 2


Based on the seismic refraction and surface wave measurements, a soil profile was established,
independent prior to the geotechnical investigations which were performed at a later stage of the
project. Figure 21 shows a generalized soil profile with a color code according to Table 4. Also
shown is the CPT in test point C1.
Cone Stress, qt (MPa)
0 5 10 15
0

10

DEPTH (m)
15

20

Fig. 21. Generalized soil profile based on seismic refraction and MASW investigations. Also
shown is CPT C1. Soil layers and wave velocities are identified in Table 4.

TABLE 4. Identification of soil layers along profile A, B and C, based on refraction and surface
wave measurements.

Seismic Units Pattern VP (m/s) VS (m/s)


1 400 <155
2 400 155 - 175
3 400 175 – 200
4 1500 200 – 225
5 1500 225 – 240
6 1500 240 – 225
7 1500 225 – 240
8 1700 > 240

4.3 Seismic Down-hole Tests


Two different types of seismic down-hole tests were performed. The SCPT tests used only one
seismic sensor, cf. Figure 5a, while the SDMT used two seismic sensors, cf. Figure 5b. There were
also differences in the method of analysis. In the case of SCPT, the first arrival time at different
levels was determined by picking peak values in the time-history trace, a method which is subject
to some uncertainty. In contrast, for the analysis of SDMT data, cross-correlation (phase shift

3rd Bolivian International Conference on Deep Foundations—Volume 2 63


method) was used to determine the time difference between wave arrivals at two separated
locations.
Figure 22 shows the determined shear wave velocities by the SCPT and SDMT method at
three locations along the test area. In general, the agreement between the two tests is good.
However, it is apparent that the S-wave velocities determined by the SCPT (one seismic sensor)
fluctuate significantly, especially at depths exceeding about 10 m.

A3 F1 G1

Fig. 22. Results of SDMT and SCPT measurements in three locations


along the test area, cf. Figure 3.

Figure 23 compares shear wave velocities measured by down-hole tests (SCPT and SDMT)
with the surface wave velocity (MASW) measurements. The general agreement between the
average values of down-hole tests and MASW tests is fair. The down-hole method provides
significantly more detailed resolution, especially in the case of the SDMT data.

5. DETERMINATION OF SOIL MODULUS


In geotechnical engineering, an important application of seismic measurements is the
determination of the deformation properties (moduli) of soils and rock. Section 3 outlines concepts
which can be used to determine, based on shear wave velocity measurements, the small-strain
shear modulus, Gmax and the large-strain (static) shear modulus, Gs, respectively. As has been
shown in the previous section, the shear wave velocity determined by the SDMT gives the most
consistent results. Therefore, the shear modulus at small strain, Gmax, and the static shear modulus
(at 0.5 % shear strain) has been determined only based on SDMT results, cf. Figure 24. The unit
weight of the different soil layers was based on the interpretation of CPUT and laboratory tests,
where available. The static shear modulus was calculated based on Eqs. (7) and (8). On average,
the modulus reduction factor, RM, varied between 0.10 (sand) and 0.20 (clayey silt).

64 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 23. Comparison of results from down-hole tests (SDMT and SCPT) and MASW
measurements in three locations along the test area, cf. Figure 17.

Fig. 24. Small-strain shear modulus, Gmax and static shear modulus, Gs (0.5 % shear strain)
determined from SDMT results in test locations A3, F1, and G1.

It should be observed that an error in the measurement of the shear wave velocity will be magnified
when calculating the shear modulus, cf. Eq. (3). For example, an error in the shear wave velocity
of 25 % will result in an error in the shear modulus of 56 %!

3rd Bolivian International Conference on Deep Foundations—Volume 2 65


Figure 25 shows the small-strain modulus, Gmax and the derived static moduli (0.5 % shear
strain), Gs and Es, respectively for the three test locations, A, F, and G. It can be observed that the
soil moduli vary both in the lateral direction (from A to G) as well as with respect to depth.

A3 F1 G1
Fig. 25. Small-strain shear modulus, Gmax and static shear modulus, Gs (0.5 % shear strain) and
Es at three test locations (A, F and G).

6. SUMMARY AND CONCLUSIONS


The geotechnical and seismic investigations at the B.E.S.T. site with relatively homogeneous soil
conditions provide a unique opportunity to compare the results of different testing methods. A
large number of tests has been carried out at well-defined investigation points.
As seismic testing is still not considered part of geotechnical routine investigations, the paper
describes the fundamentals of the three methods used at the B.E.S.T. site: seismic refraction,
surface wave (MASW), and down-hole test (SCPT and SDMT). Seismic testing methods are well-
established and described in guidance documents and standards. Of particular importance is that
the tests are performed with diligence as even minor deviations can lead to significant errors.
Another important aspect is the method of data interpretation. With the increasing computational
power and the availability of advanced analytical methods, the quality of data evaluation has
improved significantly in the recent past. An example of this development is the application of
surface wave measurements (SASW and MASW).
Probably the most important aspect of seismic testing for geotechnical applications is the
derivation of soil stiffness (modulus) based on shear wave velocity measurements. Therefore, a
significant part of the paper is devoted to the description of a practically applicable concept of
calculating the static soil modulus corresponding to a shear strain level of 0.5 %.
The results of seismic measurements are presented and compared with geotechnical
information, based on CPTU. Also, the results from different testing methods are compared. Based

66 3rd Bolivian International Conference on Deep Foundations—Volume 2


on the shear wave velocity, Vs the shear modulus at small strain, Gmax and the corresponding static
moduli Gs and E can be calculated.
From the results presented above, it can be concluded that the most reliable method of seismic
field testing is the down-hole method, using two seismic sensors. Data interpretation should be
carried out using a reliable, interpreter-independent method, such as cross-correlation (phase shift
method).

7. REFERENCES
Butcher, A. P., Campanella, R.G., Kaynia, A.M. and Massarsch, K. R., 2015. Seismic cone
downhole procedure to measure shear wave velocity - a guideline prepared by ISSMGE TC10:
Geophysical Testing in Geotechnical Engineering. XVIth International Conference on Soil
Mechanics and Geotechnical Engineering, Osaka, published in ISSMGE News 2015,
ISSMGE Bulletin, 9(2) 17-26.
Drnevich, V. P. and Massarsch, K. R. (1979). Sample Disturbance and Stress - Strain Behaviour,
ASCE Journal of the Geotechnical Engineering Division, 105(GT 9) 1001-1016.
Hardin, B. (1978). The nature of stress strain behaviour of soils. Proceedings, ASCE Specialty Conference on
Earthquake Engineering and Soil Dynamics, Pasadena, Vol. 1. pp. 3 – 30.
Massarsch, K. R. 2004. Deformation properties of fine-grained soils from seismic tests. Keynote
lecture, International Conference on Site Characterization, ISC’2, 19 – 22 Sept. 2004, Porto,
pp. 133-146.
Massarsch, K.R. 2015. Determination of shear modulus of soil from static and seismic penetration
testing. Jubilee Volume. Proceedings in honour of Prof. A. Anagnostopoulos, Technical
University of Athens, School of Civil Engineering – Geotechnical Department. Ed. M.
Kavvadas. Athens 2015. Publisher: Tsotras ISBN: 978-618-5066-30-7, pp. 335 - 352.
Park, C.B., Miller, R.D., and Xia, J., 1999, Multichannel analysis of surface waves (MASW).
Geophysics, Vol. 64, pp. 800-808.
Richart, F.E., Jr., Hall, J.R. and Woods, R.D. 1970. Vibrations of Soils and Foundations.
Englewood Cliffs, New Jersey, Prentice Hall, 414 p.
Robertson, P.K., Campanella, R.G., Gillespie, D. and Rice, A. 1986. Seismic CPT to Measure In-
Situ Shear Wave Velocity. ASCE, Journal of Geotechnical Engineering, 112(8) 791-804.
Santamarina, J. C., Klein, K.A., and Fam, M.A. 2001. Soils and Waves. John Wiley & Sons, Ltd.,
Rexdale, Ontario, Canada, 488 p.
Stokoe, K.H., Rix, G.J. and Nazarian, S. 1989. In-situ seismic testing with surface waves.
International Conference on Soil Mechanics and Foundation Engineering, Rio de Janeiro,
August 1989. Proceedings, Vol. 1. pp. 331-334.
Stokoe, K.H., II, Joh, S.H. and Woods, R.D. 2004. Geotechnical and geophysical site
characterization. International conference on site characterization, 2, ISC 2, Porto, Portugal,
19-22 September, 2004. Proceedings, Vol. 1, pp. 97-132.
Youd, T. L., Idriss, I. M., Andrus, R. D., Arango, I., Castro, G., Christian, J.T., Dobry, R., Liam
Finn, W.D., Harder Jr.L.F., Hynes, M.E., Ishihara, K., Koester, J.P., Liao, S.S.C., Marcuson
III, W.F., Martin, G.R., Mitchell, J.K., Moriwaki, Y., Power, M.S., Robertson, P.K., Seed,
R.B., and Stokoe II, K.H., 2001. Liquefaction Resistance of Soils: Summary Report from the
1996 NCEER and 1998 NCEER/NSF Workshops on Evaluation of Liquefaction Resistance
of Soils. Journal of Geotechnical and Geoenvironmental Engineering, ASCE, 127(10)
817-833.

3rd Bolivian International Conference on Deep Foundations—Volume 2 67


68 3rd Bolivian International Conference on Deep Foundations—Volume 2
Interpretation of Geotechnical Parameters using In-Situ Data for
The Bolivian Experimental Site for Testing (B.E.S.T.)

Agaiby, S.S. (1), and Mayne, P.W. (2)


(1)
Geosystems Engineering, Civil & Environmental Engineering, Georgia Institute of
Technology, Atlanta, GA, USA 30332-0355 <sagaiby3@gatech.edu>
(2)
Geosystems Engineering, Civil & Environmental Engineering, Georgia Institute of
Technology, Atlanta, GA, USA 30332-0355 <paul.mayne@ce.gatech.edu>

ABSTRACT. The engineering properties of the Bolivian Experimental Site for Testing
(B.E.S.T.) are examined in detail based on the field results of the geotechnical site investigation.
The performed investigation includes four main in-situ tests: Standard Penetration Tests (SPT),
Cone Penetration Tests (CPTu), Flat Plate Dilatometer Tests (DMT), and Downhole Shear Wave
Velocity (VsVH). Practical methods from the literature for estimating the geotechnical properties
of the site from in-situ tests are discussed with a presented comparison between the interpreted
values from the different in-situ methods. The investigated design parameters include: soil type,
unit weight (gt), effective friction angle (f’), stress history (OCR), and elastic shear modulus (E).

1. INTRODUCTION
A comprehensive geotechnical site investigation program has been carried out at the Bolivian
experimental site for testing. The testing program included both in-situ and laboratory tests to
interpret the main geotechnical parameters at the site to be used in numerous geotechnical design
applications. The in-situ testing program was comprised of 8 SPTs, 15 CPTus, 6 DMTs, and 3
VsVH measurements. Complementary laboratory tests included: grain size distribution, water
content, plastic and liquid limits tests.

2. STANDARD PENETRATION TESTS (SPT)


2.1 Overview
The standard penetration test (SPT) is performed during the advancement of a soil boring to
obtain an approximate measure of the dynamic soil resistance, as well as a disturbed drive
sample (split barrel type). The test was introduced by the Raymond Pile Company in 1902 and
remains today as one of the most common in-situ test worldwide following ASTM D 1586
standards. The main advantages of the SPT are obtaining both a sample and a number. The test is
simple, rugged, and suitable in many soil types except for soft clays and coarse gravels.

2.2 Corrections to the SPT N-value


Numerous correction factors to the measured N-value are necessary because of energy
inefficiencies and procedural variation in practice. The most important correction factor is the
energy efficiency which is obtained by a onetime calibration using the procedures outlined in
ASTM D 4633.
The efficiency of the system can be obtained by comparing the kinetic energy (KE=½mv2),
with the potential energy of the system (PE=mgh), where m = mass, v = impact velocity,

3rd Bolivian International Conference on Deep Foundations—Volume 2 69


g = the gravitational constant = 9.8 m/s2, and h = drop height. The energy ratio (ER) is defined as
KE/PE. Over the years, the standard of practice has varied from about 30% to 95% with different
hammer systems. As of 1985 when the inefficiencies were realized (e.g. Skempton 1986), the N
values corresponding to a mean ER = 60 % are the corrected values and termed N60, as given by:

𝑁!" = ER 60 ∙ 𝑁!"#$%&"' (1)

A total of eight boreholes were drilled at B.E.S.T. with SPT split-spoon samplers with
varying depths ranging from 9.5 m to 25 m. The 8 SPTs were carried out with an average
measured energy rating (ER) of 44 % covering the site location. Figures 1a and 1b present the
raw measured SPT – N values with depth in addition to the energy-corrected N60 values with
depth. The obtained samples from the drilled boreholes were used to identify the soil type at the
site using laboratory grain size distribution. Figure 1c presents the profiles of percentage of fines
(passing sieve #200) for the eight borehole locations; where "Sands" indicate a percentage of
fines less than 50% while "Clays and Silts" indicate a percentage of fines greater than 50%. The
average trend presented in Figure 1c indicates that the majority of the soil profile at B.E.S.T. can
be identified as sands with an exception to a 1-m thick clay/silt layer at a depth of 3 m and at
depth greater than 11 m.
Since SPT N-values in the same geomaterial will increase with increasing effective
overburden stress, the energy-corrected blow count (N60) is often stress-normalized to an
equivalent effective overburden stress of 1 atmosphere ≈ 100 kPa known as overburden
correction. The stress-normalized and energy-corrected blow count is referred to as (N1)60, and is
equal to:

(N! )!" = C! . N!" (2)

CN = (satm/svo')n' (3)

where CN is the stress normalization parameter, satm is atmospheric pressure in the same units as
svo' (i.e., 1 atm ≈ 1 bar ≈ 100 kPa), and n' is a stress exponent equal to 0.5 in clean sands (Liao &
Whitman, 1986; Kulhawy & Mayne 1990) and increases to 1 in clays (Mayne & Kemper, 1988).

2.3 Soil Unit Weight from SPT


When undisturbed samples or natural water contents are not available, the unit weight can be
estimated from the shear wave velocity (Vs in m/s) and depth (z in meters) as investigated by
Mayne (2001):

γ! kN m! = 8.31 log V! − 1.61 log z (4)

The relationship applies to particulate geomaterials that are not cemented or bonded, thus
would not be applicable to saprolites, rocks, cemented or structured diatomaceous or calcareous
or carbonate soils. By using the measured in-situ SPT resistance (N value), one can estimate the
corresponding shear wave velocity (Vs) value to be used in estimating in the unit weight of the
soil following Eq. 4. Figure 2a shows the interpreted soil unit weight values from the 8 SPT
soundings ranging from 16.3 to 19.4 kN m! with an average trend with depth.

70 3rd Bolivian International Conference on Deep Foundations—Volume 2


Imai and Tonouchi (1982) compiled a database of a variety of ground conditions through
Japan where they collected data points from over 400 boreholes covering different soil types
ranging from alluvial clays to diluvial clays, gravels, peats, and sands, in addition to special soils
such as loam, fill, and sirasu. The database included 1654 measured SPT resistance N values
with an average energy rating of 78 % with corresponding shear wave velocity (S-wave)
measured mainly using a suspension logging method. The direct relationship between the
measured shear wave velocity (Vs) and the SPT (N-value) can be expressed as:
!
𝑉! ( ) = 97.0 𝑁 !.!"# (5)
!

Fig. 1. B.E.S.T. SPT- N values with depth: (a) Uncorrected; (b) Corrected to 60% efficiency; and
(c) grain size distribution.

2.4 Effective Friction Angle from SPT


For sands, a developed correlation between the effective stress friction angle (f') and stress-
normalized and energy-corrected SPT resistance, (N1)60, was derived by Hatanaka and Uchida
(1996) where high quality undisturbed samples of natural sands were obtained by special
freezing method. Once mounted in the triaxial cell and allowed to thaw, specimens permitted
direct measurements of f' in triaxial compression tests. Corresponding field SPT data were
obtained at the same elevations as the undisturbed samples using a Japanese automatic trip
hammer system where energy efficiency is reported as 78 percent. For a reference 60%
efficiency in the U.S., the expression for peak f' is given as:

∅° = 15.4 N! !" + 20° (6)

In the case of clays, there is no direct relationship between the SPT resistance and the
effective stress friction angle, hence, a mean value of f' = 30 degrees is assigned (Mayne 2013).

3rd Bolivian International Conference on Deep Foundations—Volume 2 71


Figure 2b presents the interpreted effective friction angle values from the 8 SPT soundings
ranging from 27.7 to 44.5 degrees with an average trend with depth.

Fig. 2. B.E.S.T. profiles from SPT data: (a) soil unit weight (gt) and (b) effective stress
friction angle (f’).

2.5 Stress History from SPT


The preconsolidation stress (sp') is defined as the maximum effective overburden stress
experienced by the soil during its stress history. The overconsolidation ratio (OCR) is a
normalized and dimensionless parameter based on sp' and effective vertical stress (sv0'), such that:

OCR = sp'/ sv0' (7)

To overcome issues associated with laboratory methods, sp' can be evaluated using direct
correlations with in-situ test measurements such as standard penetration, cone penetration, flat
dilatometer, and/or vane shear tests that are faster, more economical, and productive than
laboratory tests. Kulhawy and Mayne (1990) investigated the relationship between the SPT
resistance (N) and the effective preconsolidation stress (sp') for 51 fine-grained soils. These were
mainly firm to stiff to hard clays which were neither sensitive nor structured, resulting in the
following expression:

𝜎! ! ≈ 0.47 ∙ 𝜎!"# ∙ N (8)

The SPT data were obtained primarily using safety hammers for which the average ER ≈
60%. Later, a more detailed study investigated the relationship between energy-corrected

72 3rd Bolivian International Conference on Deep Foundations—Volume 2


standard penetration resistance (N60) and the preconsolidation stress for different soil types as
expressed:

𝜎! ! ≈ 0.47 ∙ 𝜎!"# ∙ N!" !


(9)

where m is an exponent that depends on the soil type: m = 0.6 for clean quartzitic sands and
gravels, m = 0.7 for silty to clayey sands, m = 0.8 for sandy silts, m = 0.9 for silts to clayey silts,
and m = 1.0 for intact clays (Mayne 1992). Figure 3a presents the interpreted overconsolidation
ratios from the 8 SPT soundings ranging from 1.3 to 19 and showing an average trend with
depth.

2.6 Soil Modulus of Elasticity from SPT


Mayne & Frost (1988) investigated the results for Appalachian Piedmont residual silty to sandy
soils compiling over 160 flat dilatometer tests with supplementary routine soil borings and cone
penetrometer soundings in the vicinity of Washington DC, Virginia, and Maryland. The DMT
elastic moduli were compared with values obtained from laboratory tests and back-calculated
moduli from field performance of full-scale foundation measurements. By considering the SPT
penetration resistance measured at the same testing locations which had an average energy rating
of 60 % in the late 1980s, a direct relationship between the derived elastic modulus (E') obtained
from flat dilatometer tests (DMT) assuming v' = 0.2 and corrected SPT penetration resistance
(N60) was developed expressed by:
!.!"
E! bars = 22 ∙ 𝜎!"# ∙ N!" (10)

Figure 3b presents the interpreted soil modulus of elasticity values from the 8 SPT soundings
ranging from 10 to 280 bars and showing an overall average trend with depth.

Fig. 3. Interpreted profiles from SPT data at the B.E.S.T. site:


(a) overconsolidation ratio (OCR) and (b) soil modulus of elasticity (E).

3rd Bolivian International Conference on Deep Foundations—Volume 2 73


3. CONE PENETRATION TESTS (CPT)
3.1 Overview
The cone penetration test (CPT) involves the hydraulic pushing of an instrumented electronic
steel probe at a constant rate of 20 mm/s to obtain vertical profiles of stress, friction, and
pressure with depth. By recording continuous measurements with depth, the CPT is an excellent
tool for detailing strata changes, delineating the interfaces between soil layers, measuring soil
consistency, and detecting small lenses, inclusions, and layers within the ground. The data
presentation from a CPT sounding includes cone resistance (qt), sleeve friction (fs), and pore
water pressures (u2) plotted with depth in side-by-side graphs.
Test procedures for the CPT are performed in accordance with ASTM D 5778 using either a
standard drill rig or specialized cone truck. The advance of the probe requires the successive
addition of rods at approximately 1 m intervals. Cone penetrometer readings are taken at regular
intervals of 1-cm to 5-cm by a field computer and data acquisition system. Figure 4 presents the
raw measured 15 CPTu soundings carried out at the testing site.

Fig. 4. B.E.S.T. CPTu measurements with depth: (a) cone tip resistance, qt;
(b) sleeve friction, fs; and (c) pore water pressure, u2.

3.2 Soil Identification and Classification from CPT


For soil type identification, simple "rules of thumb" rely on one or more of the cone readings,
where a reference cone resistance value qt = 5 MPa should be identified. When the measured qt >
5 MPa, the results imply clean sands; whereas when qt < 5 MPa the readings suggest clays. For
the friction sleeve, it is convenient to plot this in terms of friction ratio, FR = fs/qt (%). As such,
clean sands are identified by FR < 1%, whereas insensitive clays exhibit FR > 4%.
In order to account for depth effects on the readings, stress-normalized CPT parameters have
been defined by Lunne, et al. (1997) as follows:

74 3rd Bolivian International Conference on Deep Foundations—Volume 2


Q = (qt-svo)/svo’ (11)

F = 100·fs/(qt-svo) (12)

To better identify the soil type, it is convenient to use the CPT material index, Ic which is
defined (Robertson & Wride, 1998):

I c = {3.47 − log(Q)}2 + {1.22 + log(F )}2 (13)

The aforementioned stress normalization for tip resistance (Q) directly with effective
overburden stress works well in soft clays and silts, however in sands the stress normalization is
proportional to the square root of effective stress, probably due to particle grain crushing or
breakage effects. In this case, a modified normalized cone tip resistance has been defined as
(Robertson, 2004; 2009):
n
(qt − σ vo ) ⎛ σ atm ⎞ (14)
Qtn = ⋅⎜ ⎟
σ atm ⎜⎝ σ vo ' ⎟⎠

where satm = 1 atmosphere ≈ 1 bar = 100 kPa and the exponent n is varying with soil type,
with typical values of 1.0 in the general case of clays (Ic > 2.95), n = 0.75 for silty soils, and n =
0.5 for clean sands (Ic < 2.05). The exponent n is a function of the material index Ic which in turn
is dependent on the modified normalized cone resistance (Q = Qtn). Therefore, an iterative
approach is needed to find the appropriate exponent n to identify the CPT material index using:

σ vo ' (15)
n = 0.381⋅ ( I c ) + 0.05 ⋅ ( ) − 0.15
σ atm

Figure 5a presents the profiles for the CPT material index, Ic with depth for the 15
conducted CPTus with an averaged trend indicating that the studied soil at B.E.S.T. can be
mainly classified as sand and/ or sand mix except for the crust layer within the top 1 m that can
be identified as gravelly sand and an intermediate 1-m thick clay layer at depth of 3 m that exists
in some soundings and several thin silty layers at depths of 6, 7, and 11 m.
A different means to classify the soil type is using empirical soil behavioral type (SBT)
charts as proposed by Robertson et al. (1986). The original 12-zone SBT system has been
updated and modified to a 9-zone classification scheme. The SBT number is determined by
plotting the CPTu data in terms of Qtn versus F. According to Robertson (2009), basic clay is
found in zone 3 while "hourglass" sands form zone 6, guidelines for the modified SBTn
classifications are identified in Table 1. Figure 5b presents the evaluated soil behavioral type
number for the 15 conducted CPTus where the soil profile mainly lies within zone 6 (sands) with
the crust in the uppermost 1 m in zone 7 (gravelly sands) and some exceptions at 3 m depth lying
in zone 3 (clays) which agrees with the previously presented classification by the CPT material
index, Ic and the grain size distribution from the samples obtained from the split-spoon samplers
along the SPTs.

3rd Bolivian International Conference on Deep Foundations—Volume 2 75


TABLE 1. Soil Behavioral Type and Zone Number as defined by CPTu Material Index, Ic

Soil SBT
Range CPT Material Index Ic
Classification Zone
Stiff sands and clays 8 and 9 (see note 1)
Sands with gravels 7 Ic < 1.31
Sands: clean to silty 6 1.31 < Ic < 2.05
Sandy mixtures 5 2.05 < Ic < 2.60
Silty mixtures 4 2.60< Ic <2.95
Clays 3 2.95< Ic <3.60
Organic soils 2 Ic > 3.60
Sensitive soils 1 (see note 2)
Notes: 1. Zone 8 (1.4 < F < 4.5 %) and Zone 9 (F > 4.5 %) and following criterion:
1
Qtn ≥
0.006 ⋅ ( F − 0.9) − 0.0004 ⋅ ( F − 0.9) 2 − 0.002
2. Sensitive soils of zone 1 identified when Q < 12 exp (-1.4 F)

Fig. 5. Soil Classification from CPT: (a) profiles of CPT material indices (Ic) and (b) soil
behavioral type number (SBTn) with depth.

3.3 Soil Unit Weight from CPT

A direct unit weight relationship with the sleeve friction has been investigated by Mayne (2014)
on a comprehensive database of sands, silts, and clays and is expressed as:

gt = gw· [1.22 + 0.15· ln(100·fs/satm+0.01)] (16)

76 3rd Bolivian International Conference on Deep Foundations—Volume 2


where gw = unit weight of water. Figure 6a shows the interpreted soil unit weight values from the
15 CPT soundings that range from 14 to 20 kN m! and showing an average trend with depth.

3.4 Effective Friction Angle from CPT


For evaluating the friction angle of sands, an elite database was compiled from special expensive
undisturbed samples of clean sands (Mayne 2006). Primarily, these sands were initially frozen
in-place using one-dimensional thermal techniques and, after careful mounting of specimens in
triaxial apparatuses with membranes and confinement, they were allowed to thaw, then sheared
to failure in triaxial compression. The results from undisturbed sands was shown to match well
with the expression derived by Kulhawy & Mayne (1990):

!! !!"#
𝜙° = 17.6° + 11.0 ∙ log (17)
! !
!!" !"#

For estimating the effective stress friction angle of clays, a typical value f' = 28° to 30° can be
assumed, or alternatively the NTH method can be adopted which involves an effective stress
limit plasticity solution for undrained penetration developed by Senneset et al. (1989) at the
Norwegian Institute of Technology (NTH). In this method, a cone resistance number (Nm) is
defined:

Nq − 1 q t − σ v0
Nm = = (18)
1 + N u ⋅ Bq σ vo ' + a'

where a' = c'· cotf' = attraction, Nq = Kp· exp[(p-2b)·tanf'] is the end-bearing factor for the cone
tip resistance, Kp = (1+sinf')/(1-sinf') is the passive stress coefficient, b = angle of plastification (-
20º < b < +20º) which defines the size of the failure zone beneath the tip, Nu = 6·tanf'·(1+tanf') is
the pore water pressure bearing factor. The full solution allows for an interpretation of a paired
set of Mohr-Coulomb strength parameters (c' and f') for all soil types.
For soft clays, it can be adopted that c' = 0, thus the term Nm reduces to the well-known
normalized cone resistance, Q = qnet/svo'. Further simplification is achieved by taking the angle b
= 0 (Terzaghi equation) for the case of undrained loading and an approximate deterministic
expression for Bq > 0.1 is obtained (Mayne 2007):

φ ' = 29.5°Bq 0.121 [0.256 + 0.336Bq + log Q] (19)

Figure 6b presents the interpreted effective friction angle values from the 15 CPT soundings
ranging from 24 to 48 degrees with an average trend with depth.

3.5 Stress History from CPT


For estimating the effective preconsolidation stress using the cone penetrometer, a general
equation for all types of natural soils, including sands, silts, clays, and mixed soil types has been
introduced by Mayne et al. (2009) with a generalized expression is expressed as:

!! !!!!
𝜎! ! = 0.33 ∙ 𝑞! − 𝜎!" ∙ 𝜎!"# 100 (20)

3rd Bolivian International Conference on Deep Foundations—Volume 2 77


where the exponent m' is a parameter that increases with fines content and decreases with mean
grain size. The approximate value of parameter m' ≈ 0.72 in clean quartz sands, 0.8 in silty sands,
up to m' = 1.0 in intact clays of low sensitivity. Using the CPT material index Ic one can identify
the magnitude of the parameter m’ for general profiling of sp' in homogeneous or heterogeneous
deposits, as well as mixed soils. For basic uncemented and non-structured soils, the exponent m’
can be estimated as follows:
!.!"
𝑚! = 1 − (21)
!! !! !.!" !"

Figure 7a presents the interpreted overconsolidation ratio values from the 15 CPT soundings
that range from 1.5 to 20 along with an overall average trend of OCR with depth.

Fig. 6. Interpreted B.E.S.T. profiles from CPT data: (a) soil unit weight (gt) and
(b) effective friction angle (φ’).

3.6 Soil Modulus of Elasticity from CPT


Elastic theory allows for interrelationships between the equivalent elastic Young's modulus (E),
shear modulus (G), and constrained modulus (D) in terms of the Poisson's ratio, such that:

E = 2 ∙ G ∙ (1 + υ) (22)

(1 − ν' )
D' = E' ⋅ (23)
(1 + ν' )(1 − 2ν' )

78 3rd Bolivian International Conference on Deep Foundations—Volume 2


For a value n’ ≈ 0.2 that is characteristic of sands and granular soils, the ratio D'/E' = 1.1 and
therefore the constrained modulus and drained Young's modulus are often used somewhat
interchangeably. For an approximate evaluation of the constrained modulus (and drained
Young's modulus) from CPT results, the common approach is expressed in the form:

D! ≈ α! ∙ q! − σ!" (24)

where aD is an empirical scaling factor that has been shown to depend upon soil type, confining
stress level, overconsolidation, and other factors (Kulhawy & Mayne, 1990). From numerous
studies in the literature, aD ≈ 5 is an approximate starting place, excepting soft plastic organic
clays and cemented geomaterials (Mayne 2007b). Figure 7b presents the interpreted soil modulus
of elasticity values from the 15 CPT soundings, with E' ranging between 5 to 60 MPa. The
overall average trend of E' is also shown with depth.

Fig. 7. Interpreted profiles from B.E.S.T. CPTs: (a) overconsolidation ratio (OCR) and
(b) soil modulus of elasticity (E).

4. FLAT PLATE DILATOMETER TEST (DMT)


4.1 Overview
The flat dilatometer test (DMT) is simple, repeatable, economic, and operator-independent. The
device consists of a tapered stainless steel blade with 18° wedge tip that is vertically advanced into
the ground at either 20-cm or 30-cm intervals per ASTM D 6635. Two pressure readings are taken
at each test depth (A and B) as a flexible steel membrane is inflated with nitrogen gas. The DMT
analyses primarily rely on empirical correlative relationships and may require adjustments for local
geologies. No borehole cuttings or spoil are generally produced by this test, although it is possible
to advance a conventional soil boring and then perform the DMT downhole within the borehole.

3rd Bolivian International Conference on Deep Foundations—Volume 2 79


4.2 DMT Index Parameters
The field A- and B- readings need to be corrected for membrane stiffness effects to obtain the
contact or lift-off pressure, po, and expansion pressure, p1. Corrections of the readings have been
presented by Schmertmann (1986):

p! = 1.05 A + ∆A − z! − 0.05 (B − ∆B − z! ) (25)

p! = B − ∆B − z! (26)

where ΔA and ΔB are reported as positive absolute values for the calibration factors for applied
suction and expansion of the membrane in air, respectively, and zm is the gage offset zero reading
when vented to atmospheric pressure (typically set to zero for a new gage). Figure 8 presents the
corrected lift-off and expansion pressure profiles with depth for the 6 DMTs performed at the
B.E.S.T. site.

Fig. 8. B.E.S.T. DMT readings with depth: (a) lift off pressure p0 and (b) expansion pressure p1.

The two dilatometer pressures, p0 and p1, are used together with the hydrostatic water
pressure, uo, to provide three index parameters: (a) material index ID, (b) horizontal stress index
KD, and (c) dilatometer modulus, ED. These were developed by Marchetti (1980) to provide
information on the stratigraphy, soil types, and the evaluation of soil parameters. Hydrostatic
water pressure (u0) can be evaluated based on available groundwater table information. The DMT
material index, ID, is related to the soil classification and is presented as:

I! = p! − p! p! − u ! (27)

80 3rd Bolivian International Conference on Deep Foundations—Volume 2


According to Marchetti (1980), the soil type can be identified: clay: 0.1 < ID < 0.6, silt: 0.6 <
ID < 1.8, and sand: 1.8 < ID < 10. In general, ID provides an expressive profile of soil type, and for
normal soils, a reasonable soil description as illustrated in Figure 9a, where by considering the
averaged trend with depth it can be determined that most of the soil profile under study can be
considered as sand mixed with sand-silt mixture with an exception for a 1-m thick clay layer at
about 3 m depth which agrees with the soil classification presented earlier by both SPTs and CPTs.
The horizontal stress index, KD, is related to the in-situ horizontal stress-state of the soil. The
index KD will always be greater than K0 due to disturbance caused during insertion of the blade.
This parameter is presented in Figure 9b and is expressed as:

K ! = p! − u ! σ′!" (28)

The dilatometer modulus ED; presented in Figure 9c; is obtained from p0 and p1 from the
theory of elasticity. For the 60-mm membrane diameter and required 1.1 mm displacement, it is
found according to Marchetti (1980) that the modulus is evaluated as:

E! = 34.7 p! − p! (29)

Fig. 9. B.E.S.T. profiles from DMTs: (a) soil material index (ID), (b) horizontal stress index (KD),
and (c) dilatometer modulus (ED).

4.3 Soil Unit Weight from DMT


The total soil unit weight (gt) can be evaluated using an approximate expression from the
material index and dilatometer modulus as:
!.! !!.!"
γ! = 1.12 γ! E! σ!"# I! (30)

3rd Bolivian International Conference on Deep Foundations—Volume 2 81


where gw = unit weight of water and σatm = atmospheric pressure. For each successive layer, the
cumulative total overburden stress (σvo) can be calculated, as this is needed for the determination
of the effective vertical overburden stress (σvo’= σvo - uo) and the evaluation of the KD parameter
(Mayne et al., 2002). Figure 10a shows the interpreted soil unit weights from the 6 DMT
soundings, ranging from 13.5 to 19.5 kN m! , and also present an average trend with depth.

4.4 Effective Friction Angle from DMT


The peak friction angle in sands can be assessed using the flat plate dilatometer test. A wedge
plasticity solution for the CPT was presented by Marchetti (1985) that was later cross-correlated
for CPT-DMT relationships by Campanella & Robertson (1991). The wedge solutions relate the
DMT lateral stress index (KD) as a function of φ’ and lateral stress state including active, at-rest
(NC), and passive conditions. The passive case provides a generally conservative evaluation of
peak friction angle and gives a good agreement with field data from different sand sites (Mayne,
2001). The expression for the passive case can be approximated by a hyperbola in the form:
!
ϕ! = 20 ° + (31)
!.!" ! !.!" !!

As for clays and silts, Ouyang & Mayne (2016) use the NTH method, as explained earlier
with the cone penetration test in Eq. 19, with the following DMT equivalent quantities:

(u2-u0)DMT = (p0 - u0) (32)

(qt-svo)DMT = 2.93·p1 - 1.93·p0 - u0 (33)

Figure 10b shows the interpreted effective friction angle values from the 6 DMT soundings
using methods for sands and clays ranging from 26 to 45 degrees with an average trend with
depth.

4.5 Stress History from DMT


Initial studies by Marchetti (1980) investigated the relationship between the overconsolidation
ratio (OCR) and the DMT horizontal stress index, KD. The correlation was based on the results of
database from only five clays that was later investigated by Mayne (1995) to include data from
24 clays ranging from intact to calcareous to fissured clays. The correlation is expressed in terms
of net contact pressure (p0 – u0):

𝜎! ! ≈ 0.5 ∙ 𝑝! − 𝑢! (34)

Figure 10c shows the interpreted OCRs from the 6 DMT soundings that range from 1.5 to
18. The overall average trend with depth is also shown and over much of the profile suggests low
OCR soils in the range of 1 to 2.

82 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 10. B.E.S.T. profiles from DMT soundings: (a) soil unit weight (gt), (b) effective stress
friction angle (f’), and (c) overconsolidation ratio (OCR).

4.6 Soil Modulus of Elasticity from DMT


The constrained modulus MDMT determined from the flat dilatometer test is the drained tangent
modulus at σ'v0 and is equivalent to the oedometer modulus obtained in the laboratory. MDMT is
evaluated using correction factor RM to dilatometer modulus ED using adjustment factor RM:

M!"# = 𝑅! . E! (35)

The equations defining RM as a function of both ID and KD are described by (Marchetti 1980)
and are given in Table 2.

TABLE 2. Equations defining modifier term RM = MDMT/ED


Condition RM Definition
For ID < 0.6 RM = 0.14 + 2.36 log KD
For ID > 3 RM = 0.5 + 2 log KD
For 0.6 < ID < 3 RM = 0.05 + 0.15·ID + (2.45 – 0.15·ID) log KD
For KD > 10 RM = 0.32 + 2.18 log KD
If RM < 0.85 Set RM = 0.85

The Young's modulus E' of the soil skeleton can be derived from MDMT using the theory of
elasticity equation:
!! !! (!!!"!)
𝐸! = (!! !!)
𝑀!"# (36)

In sands, using a representative value for Poisson's ratio n'= 0.2, then E' = 0.9 MDMT.

3rd Bolivian International Conference on Deep Foundations—Volume 2 83


5. DOWNHOLE SHEAR WAVE VELOCITY TESTS (VsVH)
5.1 Overview
A fundamental parameter of the ground is the shear wave velocity (Vs) which can be determined
either using field geophysics or laboratory tests on small specimens. Shear waves can be
measured on all geomaterials ranging from clays and silts to sands and gravels as well as in
mixed soil types, fractured to intact rocks, and artificial ground. Thus, it serves as an excellent
reference benchmark in comparing stiffness and stress state in almost all applications.
The profile of Vs is a necessary input for static and dynamic geotechnical analyses since it
directly provides the small-strain shear modulus (initial shear modulus):

Gmax = G0 = ρ t ⋅ Vs2 (37)

rt = total mass density = gsat / ga (38)

where ga is acceleration due to gravity = 9.81 m/s2.


In-situ methods for the measurement of shear wave velocity are many and can be
classified into two main categories: non-invasive and invasive methods. Non-invasive methods
include refraction survey, reflection survey, and surface wave methods, or Rayleigh waves.
Invasive methods include cased borehole methods such as: crosshole test (CHT), downhole test
(DHT), uphole test (UHT), and P-S suspension logger, as well as direct push methods: seismic
cone penetration test (SCPT) and seismic flat dilatometer test (SDMT) which are efficient
versions of the DHT mode.
For in-situ geotechnical testing, the most common shear wave mode is obtained by DHT
(ASTM D 7400) that measures a vertically-propagating horizontally-polarized shear wave
velocity, or VsVH mode. This can be carried out using a horizontal wave generating surface source
and downhole horizontally-oriented geophones positioned in either drilled-cased-grouted
boreholes, or by direct-push SDMT as carried out at B.E.S.T. Figure 11 presents the raw
downhole shear wave velocity profiles with depth carried out as a part of the SDMTs, in addition
to the corresponding small strain shear modulus profiles evaluated using Eq. 37 and 38.

5.2 Soil Unit Weight from Shear Wave Velocity


The soil unit weight can be directly evaluated from the profile of downhole shear wave velocity
with depth as previously mentioned in Eq. 4. Figure 12a shows the interpreted soil unit weight
values from the 3 VsVH soundings ranging from 16 to 19 kN m! with an average trend also
presented with depth.

84 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 11. Profiles of downhole shear wave velocity (VsVH) with depth and corresponding small
strain shear modulus (Go).

5.3 Effective Friction Angle from Shear Wave Velocity


A direct relation between stress-normalized shear wave velocity and effective friction angle for
clean quartz sands has been introduced by Uzielli et al. (2013), where a dataset of a total of 16
sands sampled from 12 sites using special undisturbed freezing techniques has been investigated
and expressed as:

f' = 3.90 (Vs1)0.44 (39)

where Vs1 = (Vs) / (svo' / satm) 0.25 = stress-normalized shear wave velocity (m/s) (40)

As for clays and silts, there is no direct developed relationship with DHT, hence, an
assumed friction value of 30 degrees can be assigned. Better yet, the procurement of undisturbed
samples for triaxial compression tests is recommended. Figure 12b shows the interpreted
effective friction angle values from the three VsVH profiles shows and interpreted f' varying from
35 to 45 degrees along with an overall average trend with depth.

5.4 Stress History from Shear Wave Velocity


A compiled database covering different geomaterials has investigated the evaluation of soil stress
history in terms of shear wave velocity. The shear wave velocity is used to evaluate the small-
strain shear modulus as per Eqns. 37 and 38, and by knowing the shear wave velocity profile
with depth the unit weight can be evaluated using Eq. 4 which is used to evaluate the effective

3rd Bolivian International Conference on Deep Foundations—Volume 2 85


vertical overburden stress. Accordingly, the effective preconsolidation stress can be evaluated
using (Mayne 2005):

σ! ! = 0.101 (σ!"# )!.!"# (G! )!.!"# (σ!" ! )!.!" (41)

where 𝜎!"# is atmospheric pressure = 1 bar = 100 kPa = 1 kg/cm2. Figure 12c shows the
interpreted overconsolidation ratio values from the three VsVH profiles. Corresponding OCRs
range from 1.5 to 6.2 with an average profile of interpreted OCR shown with depth.

Fig. 12. B.E.S.T. profiles from VsVH data: (a) interpreted soil unit weight (gt), (b) effective
friction angle (f’), and (c) overconsolidation ratio (OCR).

6. COMPARISON AND CONCLUSIONS


By considering the presented field measurements and the interpreted geotechnical parameters using
different in-situ measuring techniques, it can be concluded that there is a relatively good agreement
and consistency amongst the various tests. By considering the soil classification and stratigraphy,
we find that the grain size distribution results obtained from the SPT split-spoon samplers agree
with the classification methods by CPT; whether using the material index, Ic or the soil behavioral
type number (SBTn); that also agree with the results obtained using DMT material index, ID
indicating the overall general sandy nature of the soil profile with the existence of a one-m clay
layer at about 3 m depth and silty layers around 0.5 m thick at depths of 6, 7, and 11 m.
A comparison of the interpreted mean trends with depth for the different geotechnical
parameters using SPT, CPT, DMT, and Vs is presented in Figure 13 for soil unit weight, gt and
effective friction angle, f′ and in Figure 14 for OCR and soil modulus. Overall, it can be observed
that there is reasonable agreement from the different methods within an acceptable deviation of ±
10 %. Perhaps, an observed exception is the SPT interpreted friction angle data which may be
attributed to the low energy rating (ER) assigned to this equipment.

86 3rd Bolivian International Conference on Deep Foundations—Volume 2


Fig. 13. Comparison between averaged interpreted soil unit weight (gt) and effective friction
angle (f’) values from CPT, SPT, DMT, and VsVH data.

Fig. 14. Comparison between averaged interpreted overconsolidation ratio (OCR) and soil
modulus of elasticity (E) values from CPT, SPT, DMT, and VsVH data.

3rd Bolivian International Conference on Deep Foundations—Volume 2 87


7. REFERENCES
Campanella, R.G. and Robertson, P.K. 1991. Use and interpretation of a research dilatometer.
Canadian Geotechnical Journal 28(1): 113-126.
Hatanaka, M. and Uchida, A. 1996. Empirical correlation between penetration resistance and
effective friction of sandy soil, Soils & Foundations, 36(4): 1-9.
Imai, T., and Tonouchi, K. 1982. Correlation of N-value with S-wave velocity and shear
modulus. Proc. 2nd European Symp. on Penetration Testing, Vol. 1, Amsterdam: 67 – 72.
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. www.epri.com
Liao, S.S., and Whitman, R.V. 1986. Overburden Correction Factors for SPT in Sand. Journal of
Geotechnical Engineering, ASCE, 112 (3): 373-377.
Lunne, T., Robertson, P.K. and Powell, J.M. 1997. Cone Penetration Testing in Geotechnical
Practice. EF Spon/Blackie Academic, Routledge Publishers, London, 312 p.
Marchetti, S. 1980. In-situ tests by flat dilatometer. Journal of the Geotechnical Engineering
Division (ASCE), 106 (3): 299-321.
Marchetti, S. 1985. On the field determination of K0 in sand. Proc. 11th International Conf. on Soil
Mechanics & Foundation Engrg. (San Francisco), Vol. 5, Balkema, Rotterdam: 2667-2672.
Mayne, P.W. 1992. In-situ characterization of Piedmont residuum in Eastern US. Proc. National
Science Foundation (NSF) US-Brazil Geo-Workshop: Application of Classical Soil
Mechanics to Structured Soils, Belo Horizonte: 89–93.
Mayne, P.W. 2005. Integrated ground behavior: in-situ and lab tests, Deformation Characteristics
of Geomaterials, Vol. 2 (Proc. IS Lyon'03), Taylor & Francis Group, London: 155-177.
Mayne, P.W. 2006. In-situ test calibrations for evaluating soil parameters. Characterisation and
Engineering Properties of Natural Soils, Vol. 3, (Proc. Singapore), Taylor & Francis Group,
London: 1602-1652.
Mayne, P.W. 2006. The 2006 James K. Mitchell Lecture: Undisturbed sand strength from seismic
cone tests. Geomechanics & Geoengineering: 1(4): 239-257.
Mayne, P.W. 2007. NCHRP Synthesis 368 on Cone Penetration Test. Transportation Research
Board, National Academies Press, Washington, D.C., 118 pages. Download from:
www.trb.org
Mayne, P.W. 2007b. Invited Overview Paper: In-situ test calibrations for evaluating soil
parameters, Characterization & Engineering Properties of Natural Soils, Vol. 3 (Proc.
Singapore 2006), Taylor & Francis Group, London: 1602-1652.
Mayne, P.W. 2008. Keynote paper: Piezocone profiling of clays for maritime site investigations.
Geotechnics in Maritime Engineering, Vol. 1 (Proceedings, 11th Baltic Sea Geotechnical
Conference, Gdansk), Polish Committee on Geotechnics: 333-350.
Mayne, P.W. 2013. Updating our geotechnical curricula via a balanced approach of in-situ,
laboratory, and geophysical testing of soil. Proceedings, 61st Annual Geotechnical Conference,
Minnesota Geotechnical Society, University of Minnesota, St. Paul: 65-86.
Mayne, P.W. 2014. Generalized CPT method for evaluating yield stress in soils.
Geocharacterization for Modeling and Sustainability (GSP 234: GeoCongress, Atlanta),
ASCE, Reston, Virginia: 1336-1346.
Mayne, P.W. 2014. KN2: Interpretation of geotechnical parameters from seismic piezocone tests.
Proc. 3rd International Symposium on Cone Penetration Testing (CPT'14, Las Vegas), edited
by P.K. Robertson and K.I. Cabal: 47-73. www.cpt14.com

88 3rd Bolivian International Conference on Deep Foundations—Volume 2


Mayne, P.W. 2015. Peak friction angle of undisturbed sands using DMT. Proceedings, Third
International Conference on Flat Dilatometer, (DMT'15, Rome), paper ID 20: 237-242:
www.marchetti-dmt.it
Mayne, P.W. and Brown, D.A. 2003. Site characterization of Piedmont residuum of North
America. Characterization and Engineering Properties of Natural Soils, Vol. 2, Swets &
Zeitlinger, Lisse: 1323-1339.
Mayne, P.W. and Campanella, R.G. 2005. Versatile site characterization by seismic piezocone
tests. Proc. 16th Intl. Conf. on Soil Mechanics & Geot. Engrg, (ICSMGE, Osaka), Vol. 2,
Millpress, Rotterdam: 721-724.
Mayne, P.W. and Frost, D.D. 1988. Dilatometer Experience in Washington, DC and Vicinity,
Transportation Research Record 1169, National Academy Press, Washington, DC: 16-23
Mayne, P.W. and Kemper, J.B. 1988. Profiling OCR in Stiff Clays by CPT and SPT. ASTM
Geotechnical Testing Journal, 11(2) 139-147.
Mayne, P.W., Christopher, B., Berg, R., and DeJong, J. 2002. Subsurface Investigations -
Geotechnical Site Characterization. Publication No. FHWA-NHI-01-031, National Highway
Institute, Federal Highway Administration, Washington, D.C., 301 pages
Mayne, P.W., Coop, M.R., Springman, S., Huang, A-B., and Zornberg, J. 2009. State-of-the-Art
Paper (SOA-1): Geomaterial Behavior and Testing. Proc. 17th Intl. Conf. Soil Mechanics &
Geotech. Engrg., (Alexandria, Egypt), Vol. 4, Millpress/IOS Press Rotterdam: 2777-2872.
Ouyang, Z. and Mayne, P.W. 2016. Effective friction angle of soft-firm clays from flat dilatometer.
Geotechnical Engineering, Intl. Journal of the Institution of Civil Engineers, London: Paper
1600073. pub. online 29 Nov 2016.
Robertson, P.K. 2004. Evaluating soil liquefaction and post-earthquake deformations using the
CPT. Geotechnical & Geophysical Site Characterization, Vol. 1 (Proc. ISC-2, Porto),
Millpress, Rotterdam: 233-249.
Robertson, P.K. 2009. Interpretation of cone penetration tests: a unified approach. Canadian
Geotechnical Journal, 46(11) 1337-1355.
Robertson, P.K. and Wride, C.E. 1998. Evaluating cyclic liquefaction potential using the cone
penetration test. Canadian Geotechnical Journal, 35(3)442-459.
Robertson, P.K., Campanella, R.G., Gillespie, D. and Greig, J. 1986. Use of piezometer cone data.
Use of In-Situ Tests in Geotechnical Engineering (GSP 6), ASCE, Reston/VA: 1263-1280.
Schmertmann, J.H. 1986. Dilatometer to compute Foundation Settlement. Use of In- Situ Tests in
Geotechnical Engineering, (Proc. In-Situ'86, Virginia Tech, Blacksburg, VA), GSP 6, ASCE,
Reston, Virginia: 303-321.
Senneset, K., Sandven, R. and Janbu, N. 1989. Evaluation of soil parameters from piezocone tests.
Transportation Research Record 1235, National Academy Press, Washington DC: 24-37.
Skempton, A.W. 1986. Standard penetration test procedures and effects in sands. Geotechnique
36(3) 425-447.
Terzaghi, K., Peck, R.B., and Mesri, G. 1996. Soil Mechanics in Engineering Practice. John Wiley
& Sons: 565 p.
Uzielli, M., Mayne, P.W. and Cassidy, M.J. 2013. Probabilistic assessment of design strengths for
sands from in-situ testing data. Modern Geotechnical Design Codes of Practice, Advances in
Soil Mechanics & Geotechnical Engineering, Vol. 1, IOS-Millpress, Amsterdam: 214-227.

3rd Bolivian International Conference on Deep Foundations—Volume 2 89


90 3rd Bolivian International Conference on Deep Foundations—Volume 2
ISSMGE Bulletin: Volume 9, Issue 2 Page 17
TC Corner
TC10 – Seismic cone downhole procedure to measure shear wave
velocity: A guideline
i)
ISSMGE Technical Committee 10, “Geophysical Testing in Geotechnical Engineering”, (TC 10) was
established in 1989. The work of TC 10 was led by a Core Group, consisting of: Mr. Tony Butcher, United
Kingdom, Dr. Amir Kaynia, Norway; Dr. K. Rainer Massarsch, Sweden (Chair); Dr. Nils Rydén, Sweden
(Secretary, 2004 - 2005) & Dr. Anders Bodare (Secretary, 2001 – 2003); Prof. Kohji Tokimatsu, Japan and
Dr. Bob Whiteley, Australia.

A primary objective of TC 10 was to develop the Guidance Document “Seismic Cone Downhole Procedure
to Measure Shear Wave Velocity - A Guideline”. A Task Force was set up to implement the development of
a Guidance document on Seismic Cone Downhole Testing (SCPT). The members of the Task force were:
Tony Butcher (Chairman), Richard Campanella, Amir Kaynia and K. Rainer Massarsch. A Draft of the
document was presented at the TC 10 Member Meetings in, respectively, Prague and Porto, for the
occasion of the 2nd International Conference on Site Characterization, organized by ISSMGE. Technical
Committee on “Ground Property Characterization from In Situ Testing”, TC16, presently TC102.
Thereafter, the Final Draft was submitted to TC 10 members and sister TCs for commenting. The final
document was presented at the TC 10 Member Meeting, which was held in connection with the 16 th
ICSMGE, held in Osaka 2005. However, the document was never published in the proceedings of the Osaka
conference.

The activities of TC 10 have since been merged with those of ISSMGE TC102. Upon suggestion by ISSMGE
President Roger Frank, and with the support of TC 102 chairman Antonio Viana da Fonseca, the document
is endorsed by TC 102 and now formally published in this issue of ISSMGE Bulletin. The formal reference to
the document is:

Butcher, A. P., Campanella, R.G., Kaynia, A.M. and Massarsch, K. R., 2005. “Seismic cone downhole
procedure to measure shear wave velocity - a guideline”, prepared by ISSMGE TC10: Geophysical Testing
in Geotechnical Engineering. ISSMGE Bulletin April 2015 issue.

3rd Bolivian International Conference on Deep Foundations—Volume 2 91


ISSMGE Bulletin: Volume 9, Issue 2 Page 18
TC Corner
TC10 – Seismic cone downhole procedure to measure shear wave
velocity: A guideline (Con’t)
ii)
Seismic cone downhole procedure to measure shear wave velocity - a guideline
prepared by ISSMGE TC10: Geophysical Testing in Geotechnical Engineering
Procédé séismique de downhole de cône à la vitesse d’ondes de cisaillement de mesure - une directive a
préparé par ISSMGE TC10 : Essai géophysique dans la technologie géotechnique

A.P. Butcher (BRE, UK), R.G. Campanella (University of British Columbia, Canada), A.M. Kaynia
(Norwegian Geotechnical Institute, Norway) and K.R. Massarsch (Geo Engineering AB, Sweden)

Abstract
The International Society for Soil Mechanics and Geotechnical Engineering, Technical Committee No. 10:
Geophysical Testing in Geotechnical Engineering has as part of its brief the task of drafting guidelines for
geophysical techniques where no other national or international standards or codes of practice exist. This
document is the first of these guidelines and concerns the use of the Seismic Cone to measure downhole
seismic wave propagation.

Resume
La Société Internationale de Mécanique des Sols et de la Géotechnique, le Comité technique No. 10:
L'essai géophysique dans la technologie géotechnique a en tant qu'élément de son dossier le charger des
directives de rédaction pour des techniques géophysiques où aucune autre norme ou recueil d'instructions
nationale ou internationale n'existe. Ce document est le premier de ces directives et concerne l'utilisation
du cône séismique de mesurer la propagation séismique d’ondes de downhole.

Introduction
This document is to provide guidance to practitioners and procurers on downhole seismic wave
measurement using a seismic cone penetrometer. The guideline has been prepared by ISSMGE TC10:
Geophysical Testing in Geotechnical Engineering and is a supplement to the International Reference Test
Procedure (IRTP) for the electric Cone Penetration Test (CPT) and the Cone Penetration Test with Pore
pressure (CPTU) as produced by the ISSMGE TC16. The document therefore follows, and should be used
with, the CPT IRTP (1999).

The addition of a seismic sensor (usually a geophone but may be an accelerometer or seismometer) inside
the barrel of a standard electric CPT is termed a Seismic Cone Penetrometer Test (SCPT) (Robertson et al.,
1986). Such a sensor allows the measurement of the arrival of vertically propagating seismic body waves,
generated from a source on the ground surface, in addition to the usual cone parameters that are used for
detailed stratigraphic logging.

There are two types of seismic body waves, Pressure or Compression waves (P waves) as well as Shear
waves (S waves) and seismic sensors react to both. The P wave always arrives first. In soils below the
ground water table the P wave typically travels 2 or more times faster than the S wave, so separation of
the two body waves is easy. Above the water table, however, the difference is small and separation of P
and S waves may be very difficult, requiring specialized techniques. However the most significant
difference between P and S waves is that S waves are reversible. Therefore using a source that can
produce shear waves of opposite polarity facilitates the identification of S waves.

Since shear waves travel through the skeletal structure of the soil at very small strains, one can apply
simple elastic theory to calculate the average elastic small strain shear modulus, over the length interval
of measurement, as the mass density times the square of the shear wave velocity. Thus, the shear wave
velocity relates directly to stiffness (Massarsch, 2004) and may also be used to estimate liquefaction
susceptibility in young uncemented sands (Youd et al., 2001).

92 3rd Bolivian International Conference on Deep Foundations—Volume 2


ISSMGE Bulletin: Volume 9, Issue 2 Page 19
TC Corner
TC10 – Seismic cone downhole procedure to measure shear wave
velocity: A guideline (Con’t)
iii)
Definitions
The following definitions will be used:
 Accelerometer: Sensor that produces an output in response to a seismic wave by way of a change in
capacitance caused by the relative movement of a mass and the sensor case. An accelerometer
detects particle accelerations.
 Array: group of devices at one location orientated orthogonally to each other.
 Data recording equipment: Equipment to log the signals from the seismometers.
 Geophone: Sensor that gives an output in response to seismic waves using the relative movement of a
mass (magnet) moving within a coil fixed to the sensor case. A geophone detects particle velocities.
 Hammer: Heavy mass to impact the Shear Beam as part of the Source.
 Interval time: The difference in arrival times of seismic waves at the receivers at two
depths/distances from the Source. The ‘true interval’ is the difference in arrival times between
receivers at a fixed distance apart and the ‘pseudo interval’ is the difference in arrival times to the
same receiver when placed at two different distances from the source.
 Seismometer: Device that produces a calibrated self generated output response to imposed seismic
waves and gives maximum output at its natural frequency or fundamental mode (goes into resonance)
when activated by seismic waves. A seismometer can be an accelerometer, geophone or a sensor able
to detect deflections in the range 0 to 250 Hz.
 Seismometer natural frequency: Frequency at which the seismometer gives its maximum output and
above which the seismometer response is constant.
 Shear beam: Beam that forms part of the downhole seismic shear wave Source that is impacted by a
Hammer to maximize S waves and minimize P waves.
 Source: Device that, when activated, generates polarised shear waves that propagate into the ground.
(A basic source will include a loaded Shear Beam, Hammer and a Trigger to activate the data
recording equipment).
 Trigger: Device attached to either the Shear Beam or the Hammer to initiate the data recording
equipment at the instant the Shear Beam is struck by the Hammer.

Methodology
During a pause in cone penetration, a shear wave can be created at the ground surface that will propagate
into the ground on a hemi-spherical front and a measurement made of the time taken for the seismic
wave to propagate to the seismometer in the cone. By repeating this measurement at another depth, one
can determine, from the signal traces, the interval time and so calculate the average shear wave velocity
over the depth interval between the seismometers. A repetition of this procedure with cone advancement
yields a vertical profile of vertically propagating shear wave velocity. Fig. 1 shows two alternative
schematic arrangements of the SCPT, and Fig. 2 shows a typical arrangement of the surface shear wave
source.

Equipment
The general arrangement of equipment is shown in Figs 1 and 2.

Seismometer: The seismometer will typically have a natural frequency of less than 28 Hz and must fit
inside the cone barrel. The seismometer must be mounted firmly in the cone barrel with the active axis in
the horizontal direction and the axis alignment indicated on the outside of cone body. The cone barrel at
the location of the seismometer should be of a greater diameter than the sections immediately below the
location of the seismometer to ensure good acoustic coupling between the cone barrel and the
surrounding soil.

3rd Bolivian International Conference on Deep Foundations—Volume 2 93


ISSMGE Bulletin: Volume 9, Issue 2 Page 20
TC Corner
TC10 – Seismic cone downhole procedure to measure shear wave
velocity: A guideline (Con’t)
iv)
Axis of SCPT Axis of SCPT
Shear Shear
beam X beam
X

CPT push CPT push


rods rods
D1
D1 L1
L1
Assumed travel pat hs of Assumed travel pat hs
seismic waves from shear of seismic waves from
beam to seismometers in shear beam to
SCPT at depth D1 SCPT body at depths D1 Receiver 1 L2 seismometers in
L2
and D2 SCPT body

D2 Dual D2
array
SCPT
body
SCPT at depth D2 Receiver 2

Figure 1a. Schematic diagram of the seismic cone Figure 1b. Schematic diagram of the dual array
test with required dimensions, D1, D2, and X seismic cone test with required dimensions, D 1, D2,
and X

Comment: Some seismic cones include 2 seismometers in


an array in the horizontal plane set with their active
axes orthogonally. This configuration allows
compensation for possible rotation of the cone drive
rods, (and the cone containing the seismometer) with
the subsequent loss in response and also gives
orthogonal seismic wave traces from the same source
activation. In variable and layered ground conditions, Shear
beam
with ambient noise or ground structures that would
corrupt the received signals, wave characteristics of the
source can be used to identify the shear wave amongst
the other waves.

The inclusion of a vertically orientated seismometer will


allow the P wave element of the seismic wave to be Shear
Distance (X) from
assessed or P wave arrival measured if a P wave source beam
the centre line of
shear beam to
is used. In many cases the combination of P and S wave insertion point of

data can enhance the identification of stratigraphic Insertion


point of
seismic cone

boundaries. seismic
cone

Figure 2. Typical downhole shear wave


source setup with shear beam and fixed axis
swing hammers.

94 3rd Bolivian International Conference on Deep Foundations—Volume 2


ISSMGE Bulletin: Volume 9, Issue 2 Page 21
TC Corner
TC10 – Seismic cone downhole procedure to measure shear wave
velocity: A guideline (Con’t)
v)
Shear Beam: The beam can be metal or wood encased at the ends and bottom with minimum 25 mm thick
steel. The strike plates or anvils at the ends are welded to the bottom plate and the bottom plate should
have cleats welded to it, to penetrate the ground and prevent sliding when struck. The shear beam is
placed on the ground and loaded by the levelling jacks of the cone pushing equipment or the axle load
from vehicle wheels. The ground should be prepared to give good continuous contact along the whole
length of the beam to ensure good acoustic coupling between the beam and the ground. The Shear Beam
should not move when struck by the hammers otherwise energy is dissipated and does not travel into
ground and does not produce repeatable seismic shear waves. The anvils, on the ends of the Shear Beam,
when struck in the direction of the long axis of the Shear Beam, will produce shear waves of opposite
polarity.
Comment: The beam can be continuous (approximately 2.4 m long) i.e. greater than the width of a
vehicle or equipment used to load the beam and 150 mm wide or alternatively can be two shorter beams
placed and loaded so that the anvils oppose and can be struck by the hammers to produce shear waves of
opposite polarity. Care must be taken to position the beams and strike direction to maximise S waves and
minimise the production of P waves.

Heavy hammer(s): Heavy hammer(s) with head mass of between 5 to 15 kg to strike the plate or anvil on
the end of the shear beam in a direction parallel to the long axis of the shear beam and the active axis of
seismometer.
Comment: Two fixed axis hammers, one to strike each end of the beam in the specified directions, will
significantly speed up the operation and give controllable and consistent source output. A typical setup
is shown in Fig. 2.

Data recording equipment: The recording equipment can be a digital oscilloscope, a P.C. with installed
A/D board and oscilloscope software or a commercial data acquisition system such as a seismograph. The
data recording equipment must be able to record at 50 s (microsecond) per point interval, or faster, to
ensure clear uncorrupted signals and to start the logging of the seismometer outputs using an automatic
trigger. An analogue anti-aliasing filter should be used to avoid corruption of signal frequencies above the
device limits. Commercial data recording equipment usually include amplifiers and signal filters to help
enhance recorded signals. The effect of these processes on the recorded signals must be considered
before their use. For example, filtering can cause phase shift of signals and amplification is usually limited
to a frequency range. In either case, the signals may not be directly comparable.
Comment: Experience has shown that there is a significant advantage to record the unprocessed data and
then the effect of filtering and processing can be assessed during post processing. Most modern
acquisition equipment allows the viewing of filtered signals during acquisition (to assess quality and
repeatability) but saves the data un-filtered. Most modern acquisition equipment allows signal stacking
to improve signal to noise ratio.

Trigger: The trigger can be fixed to the hammer head or the beam. The trigger is required to be very fast
(less than 10 microsecond reaction time) and repeatable. When the hammer hits the shear beam, the
electrical reaction of the trigger activates the trigger circuit that outputs to the signal recording
equipment. A typical trigger circuit is given in Campanella & Stewart (1992). A seismic trigger mounted
on the beam may be used if it is fast enough, repeatable and delay time is checked and known or a
contact trigger that works the instant contact is made between the hammer and the anvil.
Comment: The use of 2 arrays of seismometers set in the cone barrel a fixed distance apart, say 0.5 m or
1.0 m, (termed a dual array seismic cone, see Fig. 1b) would enable the travel time of the shear wave to
be measured between the seismometers from the same source activation thereby avoiding possible errors
from selection of signal from different source activation, the speed of the trigger, and the accuracy of
distance from the source to the receivers from successive pushes of the drive rods to each depth. In this
case the seismometers must have identical response characteristics (natural frequency, calibration and
damping). However if signals are to be stacked, that is the signals from successive source activations
added together to improve signal to noise ratio, the trigger time must be repeatable.

3rd Bolivian International Conference on Deep Foundations—Volume 2 95


ISSMGE Bulletin: Volume 9, Issue 2 Page 22
TC Corner
TC10 – Seismic cone downhole procedure to measure shear wave
velocity: A guideline (Con’t)
vi)
Test procedures
At the start of the SCPT, the body of the cone should be rotated until the axis of a seismometer is parallel
to the long axis of the shear beam.

(a) The cone is pushed into the ground, monitoring the inclination of the cone barrel during the push.

Comment: It is important to know the exact location of the receivers in all three axes and the
inclinometer in the cone barrel will give the horizontal component and the depth measuring system of
the CPT the vertical component.

(b) The penetration of the cone is stopped and the seismometer depth is recorded. The horizontal offset
distance, X, from cone to centre of the shear beam should also be recorded (see Fig. 1).

Comment: Typically this procedure is carried out at depths greater than about 2-3 m in order to minimize
the interference of surface wave effects. If the seismic cone includes a fully operative electric cone then
it will be advanced at 2 cm/s and stopped typically at a rod break at 1m intervals or for pore water
pressure dissipation tests. If acceptable, such stoppages can also be used for downhole seismic wave
measurements. Alternatively the seismic cone can be pushed to a predetermined depth at which the
shear wave velocities are required and the measurements made. To avoid the possible effects of time
between stopping, pushing and making measurements, it is advisable to keep this time interval consistent.
The horizontal distance, X, between the entry point of the seismic cone and the source should be kept at
around 1m. Greater distances will require the effects of curved travel paths, that particularly affect
single array SCPT’s, to be addressed. It is advisable at the first depth of measurement to monitor the
output of the receivers without activating the source to determine the ambient seismic noise in the
ground and thereby enable the filtering, as far as possible, the ambient noise. Experience has shown that
ambient noise can be reduced by retracting the cone pushing system, so that the drive rods are unloaded
and there is no contact between the shear beam system and the cone drive rods through the cone drive
vehicle, and the cone driving equipment motors are not running.

(c) The shear beam is struck by the hammer and the trigger activates the recording equipment that then
displays the time based signal trace received by the seismometer.

Comment: For quality assurance, it is recommended to reset the trigger and repeat the procedure until a
consistent and reproducible trace is obtained. The voltage-time traces should lie one over the other. If
they do not, continue repeating until measured responses are identical. In the case of the dual array
SCPT the traces from both the seismometers can be displayed together giving a rapid assessment of the
shear wave propagation time. If the seismic wave velocity appears too high then there may be a
connection between the cone drive system and the seismic cone so allowing the seismic waves to travel
through the cone drive rods instead of the ground.

(d) The trigger is reset and the shear beam is then struck by the hammer on the opposite end on the
other side of vehicle (causing initial particle motion in the opposite direction and a shear wave of
opposite polarity) and procedure in step (c)) is again completed.

(e) Show the traces from steps (c) and (d) together and identify the shear wave (usually clearly seen with
traces from the opposite polarity shear waves as a mirror image in time) and pick an arrival time. An
example of a pair of signals is shown in Fig. 3.

96 3rd Bolivian International Conference on Deep Foundations—Volume 2


ISSMGE Bulletin: Volume 9, Issue 2 Page 23
TC Corner
TC10 – Seismic cone downhole procedure to measure shear wave
velocity: A guideline (Con’t)
vii)
With reversed image traces, the first major cross-over can be taken as the “reference” arrival, or one
trace may be used and an arrival pick made visually by an experienced operator. If the wave arrival point
is not clear then a significant point that occurs on both traces can be used provided it occurs shortly after
the likely wave arrival, later selections are likely to be affected by signal attenuation and dispersion.
Alternately, a cross-correlation procedure may be used to find the interval travel time using the wave
traces from strikes on the same side at successive depths (Campanella & Stewart, 1992). This technique is
more complex, but eliminates the arbitrary visual pick of arrival time and is necessary if symmetry of
reverse wave traces is lacking. If a dual array seismic cone is used then the wave traces from each
seismometer can be compared to get the travel time between seismometers. Fig. 4 shows an example of
‘pseudo interval’ traces between 4 and 15 m depth.

Comment: As depth increases the signal to noise ratio decreases. At large depths it may be necessary to
increase signal/noise (depending on the amplification, resolution and accuracy of the data recording
equipment). This can be achieved by using multiple source activation events (from 4 to 10) and adding
(or stacking) the measured signals. This will reduce most of the random noise and increase signal/noise
ratio.

Time (milliseconds)
0 40 80 120 160
0

t2-t1= 5.53ms
T2 -T1 = 5.53ms
Depth (m)

10

15

Figure 3. An example of oppositely polarised shear


Figure 4. Example of ‘pseudo interval’ traces of
wave traces with clear crossover of traces showing
shear waves at depths 4m to 15m
the interval time T2 – T1

3rd Bolivian International Conference on Deep Foundations—Volume 2 97


ISSMGE Bulletin: Volume 9, Issue 2 Page 24
TC Corner
TC10 – Seismic cone downhole procedure to measure shear wave
velocity: A guideline (Con’t)
viii)
The average downhole shear wave velocity is calculated for the depth interval the cone has been driven
between measurements or the fixed distance between the two seismometer sets in a dual array seismic
cone. L -L
V = 2 1
S
T2 - T1
The average shear wave velocity for the given depth interval in units of m/s and assuming straight ray
paths (see Fig. 1) is given by Equation (1):

(1)
where
L1 = calculated length, m of the straight travel path distance from source to receiver at shallower depth
(use horizontal offset, X, and vertical depth D 1).
L2 = calculated length, m of the straight travel path distance from source to receiver at greater depth (use
horizontal offset, X, and vertical depth D2).
T1 = shear wave travel time from source to receiver at shallower depth (along wave path L 1).
T2 = shear wave travel time from source to receiver at greater depth (along wave path L 2).
T2 -T1 = interval travel time.

Reporting of results and interpretation procedures


The following information shall be reported:

For each site:


(a) Length of shear beam (lengths if two beams are used) and material and composition including anvils
(b) Mass of swing hammers
(c) Fixed or free pivot point of swing hammers
(d) Trigger type and location. (for single seismometer seismic cones a typical trigger delay time)
(e) Distance (X) of shear beam from insertion point of SCPT, and distance of impact points from the
insertion point of the SCPT
(f) Type of receivers, their specifications, serial numbers and name of manufacturer and last dated
response calibration
(g) Type, serial number and specification of data recording equipment and name of manufacturer

For each location:


(h) Date and time of test
(i) Identification of test
(j) Altitude and location of insertion point of SCPT

For each depth:


(k) Depth of receiver(s) from ground level
(l) Direction of swing hammer action
(m) Rate of sampling and sample length for each record.
(n) Name of files where raw and processed data are recorded including media and location of storage
(o) Type and specification of real time processing included in the recorded data
(p) Type and specification of post measurement processing included in the presented data
(q) Calculated propagation times of the shear waves and the depth range over which the measurement
was taken
(r) Calculation of the Shear Wave velocities and the depth range over which the velocity was calculated

The data files in n) should be stored for future access or for further processing until the end of the project
or as specified by the client.

98 3rd Bolivian International Conference on Deep Foundations—Volume 2


ISSMGE Bulletin: Volume 9, Issue 2 Page 25
TC Corner
TC10 – Seismic cone downhole procedure to measure shear wave
velocity: A guideline (Con’t)
ix)
Acknowledgements
Drafts of this document were discussed at the TC 10 Members Meetings in Prague (2003) and Porto (2004).
Valuable comments and suggestions for improvements were made by members of TC 10 as well as
members of TC 16 ‘Ground Properties from In-situ Testing’ and TC 1 ‘Offshore and Nearshore Geotechnical
Engineering’. Their contributions are acknowledged with gratitude.

References and further reading


Butcher, A.P. and Powell, J.J.M., 1995. Practical considerations for field geophysical techniques used to
assess ground stiffness. Proc. Int. Conf. on Advances in Site Investigation Practice, ICE London, March
1995. Thomas Telford, pp. 701-714.
Campanella, R.G. and Stewart, W.P. 1992. Seismic Cone Analysis using digital signal processing for
dynamic site characterization. Canadian Geotechnical Journal, Vol. 29, No. 3, June 1992, pp.477-486.
IRTP, 1999:ISSMGE Technical Committee TC16 Ground Property Characterisation from In-situ Testing, 1999.
International Reference Test Procedure (IRTP) for the Cone Penetration Test (CPT) and the Cone
Penetration Test with pore pressure (CPTU). Proc. XII ECSMGE Amsterdam. Balkema. pp 2195-2222.
Massarsch, K. R. 2004. Deformation properties of fine-grained soils from seismic tests. Keynote lecture,
International Conference on Site Characterization, ISC’2, 19 – 22 Sept. 2004, Porto, pp. 133-146.
Robertson, P.K., Campanella, R.G., Gillespie, D. and Rice, A. 1986. Seismic CPT to Measure In-Situ Shear
Wave Velocity. ASCE, Journal of Geotechnical Engineering, Vol. 112, No. 8, August 1986, pp. 791-804.
Youd, T. L., Idriss, I. M., Andrus, R. D., Arango, I., Castro, G., Christian, J.T., Dobry, R., Liam Finn, W.D.,
Harder Jr.L.F., Hynes, M.E., Ishihara, K., Koester, J.P., Liao, S.S.C., Marcuson III, W.F., Martin, G.R.,
Mitchell, J.K., Moriwaki, Y., Power, M.S., Robertson, P.K., Seed, R.B., and Stokoe II, K.H., 2001.
Liquefaction Resistance of Soils: Summary Report from the 1996 NCEER and 1998 NCEER/NSF Workshops
on Evaluation of Liquefaction Resistance of Soils. Journal of Geotechnical and Geoenvironmental
Engineering, ASCE, Vol. 127, No. 10, pp. 817-833.

Appendix
Maintenance, Checks and Calibrations:

This appendix contains informative guidance on maintenance, checks and calibrations for the SCPT but
excludes those parts that are common to the CPT and are included in the CPT IRTP (1999).

1. Seismometers
The seismometers should be checked to ensure they comply to the manufacturers specification in
response to seismic waves in regard to frequency, phase and damping before each profile. Where arrays of
seismometers are used, such as for true interval time measurements, each seismometer must have an
identical response, in laboratory test conditions, to seismic waves in regard to frequency, phase and
damping.

2. Source and Triggers


Where single seismometer seismic cones are used the source activation and trigger time delay will have to
be quantified. The trigger delay time needs to be repeatable and not vary by more than 1%.

3rd Bolivian International Conference on Deep Foundations—Volume 2 99


100 3rd Bolivian International Conference on Deep Foundations—Volume 2
COVER 4 COVER 1

3 BOLIVIAN
rd

3rd BOLIVIAN INTERNATIONAL CONFERENCE ON DEEP FOUNDATIONS


INTERNATIONAL
CONFERENCE ON
DEEP FOUNDATIONS

April 27 – 29, 2017


Santa Cruz de la Sierra, Bolivia

PROCEEDINGS
VOLUME 1
VOLUME 1

64427 ICDF Vol 1 Conference Cover_rb.indd 1 4/4/2017 10:06:35 AM

You might also like