Anda di halaman 1dari 546

Proceedings of

The 2nd Japan-Greece Workshop on


Seismic Design, Observation,
and Retrofit of Foundations
April 3~4, 2007, Tokyo, Japan

Proceedings of the 2nd Japan-Greece Workshop on Seismic Design, Observation, and


Retrofit of Foundations
Tokyo, Japan, April 3~4, 2007

Sponsored by
Earthquake Engineering Committee of the Japanese Society of Civil Engineers
Hellenic Society for Earthquake Engineering
Co-sponsored by
Japanese Society of Civil Engineering
Published by
Japan Society of Civil Engineers

@2007 Japan Society of Civil Engineers

All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means
electronic or mechanical, including photocopying, recording or any information storages and
retrieval system now known or to be invented, without written permission from the publisher.
978-4-8106-0620-1
Printed by WACO Co., Ltd.

CONTENTS
Organizing committee ............................................................................................................................................... i

CONFERENCE PAPERS
Tunnels, Retaining Systems, and Landslides
Deep immersed tunnel under combined major fault rupture deformation and subsequent strong seismic
shaking

Gazetas, G., Anastasopoulos, I., Gerolymos, N., Drosos, V., Kourkoulis, R. and
Georgarakos, .--------------------------------------------------------------------------------------- 1
Seismic triggering, evolution, deposition, and retaining of rapid landslides

Gerolymos, N., Gazetas G. and Vardoulakis, I. ------------------------------------------------- 27


Interaction of earthquake-triggered landslide with foundation-structure systems

Kourkoulis, R., Gelagoti, F., Anastassopoulos, I. and Gazetas, G. --------------------------- 51


Detrimental effects of urban tunnels on design seismic ground motions

Kouretzis, G., Bouckovalas, G., Sofianos, A. and Yiouta-Mitra, P. --------------------------- 64


Seismic analysis of near-surface tunnel cross-sections

Vrettos, C. -------------------------------------------------------------------------------------------- 72
Sliding of rigid block on sloping plane: the surprising role of the sequence of long-duration pulses

Garini, E. and Gazetas, G. ------------------------------------------------------------------------- 79


A replacement to the MononobeOkabe equations by stress limit analysis

Mylonakis, G., Kloukinas, P. and Papantonopoulos, C.-------------------------------------- 105


A study on ground displacement due to fault slip and its influence to underground structure

Adachi, Y., Yoshimura, S. and Nakata, T. ------------------------------------------------------ 123

Shallow Foundations and SoilStructure Interaction


Study on applicability of seismic-isolation foundation to railway structures

Luo, X., Miyamoto, T. and Imamura, T. -------------------------------------------------------- 131


A study on dynamic soil-structure interaction effect based on microtremor measurement of building and
surrounding ground surface

Iiba, M., Watakabe, M., Fujii, A., Koyama, S., Sakai, S. and Yasui, M. -------------------- 135
Damage index evaluations of SDOF system with dynamic soil-structure interaction

Kawano, K., Kimura, Y. and Nakamura, Y. ---------------------------------------------------- 143


Seismic response of rigid structures stepping on nonlinear foundation

Palmeri, A. Makris, N. ---------------------------------------------------------------------------- 150


Numerical simulations of shaking table experiments on a shallow foundation test model at PWRI, Japan

Paolucci, R., Shirato, M. and Yilmaz, M. T.---------------------------------------------------- 158

Analytical modelling of footings under large overturning moment

Apostolou, M. and Gazetas, G. ------------------------------------------------------------------ 165


Natural period and effective damping of simple structures on footings and piles

Maravas, G. Mylonakis and D. L. Karabalis -------------------------------------------------- 185


Interaction of shallow and deep foundations with a rupturing normal fault

Anastasopoulos, L. Gazetas, G.------------------------------------------------------------------ 197


The role of the soil stiffness on the dynamic impedance functions

Pitilakis, D., Clouteau, D. and Modaressi, A. ------------------------------------------------- 212


Fundamental period of sdof systems including soil-structure interaction and soil improvement

Kirtas, E., Trevlopoulos, K. , Rovithis, E. and Pitilakis, K. ---------------------------------- 225


Analytical study on seismic design of footings subjected to earthquake loading

Kosa, K., Ando, T., Adachi, Y. and Shirato, M. ------------------------------------------------ 238


Development of the sheet-pile foundation; a series of seismic proving tests

Higuchi, S., Nishioka, H., Tanaka, K., Koda, M. and Hirao, J. ----------------------------- 246
Rocking seismic isolation of bridges supported by spread foundations

Kawashima, K., Nagai, T. and Sakellaraki, D. ------------------------------------------------ 254

Deep Foundations
Effects of limit state performance of steel bearings on a bridge upper structure

Ohtomo, K., Sato, Y. and Sakai, M. ------------------------------------------------------------- 267


Large-scale model tests of shallow foundations subjected to earthquake loads

Shirato, M., Kouno, T., Nakatani, S. and Paolucci, R.---------------------------------------- 275


Numerical modeling of centrifuge cyclic lateral pile load experiments

Gerolymos, N., Drosos, V., Escoffier, S., Gazetas, G. and Garnier, J. --------------------- 299
Seismic retrofitting of the piled foundation of a reinforced concrete building

Maugeri, M. and Castelli, F. --------------------------------------------------------------------- 320


Influence of batter piles on the seismic response of pile groups

Giannakou, A., Gerolymos, N. and Gazetas, G. ---------------------------------------------- 335


Effect of the natural frequencies of the coupled soil-pile-structure system on pile dynamic response

Rovithis, E., Kirtas, E. and Pitilakis, K. -------------------------------------------------------- 344


Utilization of P-Y curves for estimating soil pile interaction under seismic loading

Rovithis1, E., Kirtas1, E. and Pitilakis1, K.---------------------------------------------------- 357


Linear elastic transient response of deep rigid foundations: From 3D finite element simulations to design

Varun, Assimaki, D. and Gazetas, G. ----------------------------------------------------------- 369


Seismic zonation, vulnerability assessment and loss scenarios in Thessaloniki

Pitilakis, K., Anastasiadis, A., Kakderi, K., Argyroudi, S. and Alexoudi, M. -------------- 386
Minimization of fixed-head pile bending at optimal radius under realistic conditions by using seismic
deformation method

Saitoh, M. ------------------------------------------------------------------------------------------- 401

Seismic response analysis of pile foundation using finite element method

Maki, T., Tsuchiya, S., Watanabe, T. and Maekawa, K. -------------------------------------- 409


Seismic behavior of single pile in cohesive soil

Tuladhar, R., Mutsuyoshi, H. and Maki, T. ---------------------------------------------------- 417


Passive and active stress wave techniques for damage assessment of railway piers

Shiotani, T., Miwa, S., Aggelis, D. G., Luo, X. and Haya, H. -------------------------------- 424

Liquefaction Densification Related Studies, and others


Centrifuge tests on remedial measure using batter piles against liquefaction-induced soil flow after quay
wall failure

Tazoh, T., Sato, M., Jang, J. and Gazetas, G.-------------------------------------------------- 431


Why do pile-supported bridge piers and not abutments collapse in liquefiable soils during earthquakes?

Kerciku, A. A., Bhattacharya, S. and H. J. Burd----------------------------------------------- 440


Effect of excess pore water pressure buildup on building damage

Dakoulas, P.---------------------------------------------------------------------------------------- 453


Numerical analysis of gravel drain performance in liquefiable soils

Papadimitriou, G., Moutsopoulou, M.-E. and Bouckovalas, G. D. ------------------------- 467


Estimating earthquake induced settlements on granular soils Application to shallow foundations

Egglezos, D. ---------------------------------------------------------------------------------------- 479


Unstable behaviour of a fine sand under cyclic loading

Georgiannou, V. N. and Tsomokos, A. ---------------------------------------------------------- 491


Numerical simulation of liquefied sand using a shear-thinning fluid model

Schenkengel, K.-U., Becker, A. and Vrettos, C.------------------------------------------------ 497


On the liquefaction resistance of silty sands

Papadopoulou, A. I. and Tika, T. M. ------------------------------------------------------------ 504


Effect of countermeasures for pile foundation under lateral flow caused by liquefaction

Matsuda, T. and Sato, K. ------------------------------------------------------------------------- 513


Effect of duration of earthquake on onset of liquefaction

Yoshida, N. ----------------------------------------------------------------------------------------- 521


A simple evaluation method for earthquake damage to the quay walls

Soejima, M., Suizu, A., Ejiri, J. and Matsuda, T. ---------------------------------------------- 527


Effect on records of seismic intensity-meter by adjacent building

Kataoka, S. ----------------------------------------------------------------------------------------- 531

Organizing Committee
Chairman

T. Katayama, Japan
G. Gazetas, Greece

Members

H. Iemura, Japan
Y. Goto, Japan
F. Miura, Japan
K. Wakamatsu, Japan
T. Tazoh, Japan
N. Yoshida, Japan
T. Matsuda, Japan
H. Kiku, Japan
M. Saitoh, Japan
K. Pitilakis, Greece
G. Bouckovalas, Greece
G. Mylonakis, Greece
A. Nikolaou, Greece
P. Dakoulas, Greece
D. Assimaki, Greece
N. Gerolymos, Greece

Tunnels, Retaining Systems, and Landslides

Deep Immersed Tunnel under Combined Major Fault Rupture


Deformation and Subsequent Strong Seismic Shaking
G. Gazetas, I. Anastasopoulos, N. Gerolymos, V. Drosos,
R. Kourkoulis, . Georgarakos
National Technical University of Athens, Greece

Abstract
Immersed tunnels are particularly sensitive to tensile deformations such as those imposed by an
earthquake normal fault rupturing underneath, and those generated by the dynamic response
due to seismic shaking. The paper investigates the response of a future 70 m deep immersed
tunnel to the combined action of a major normal fault rupture due to an earthquake occurring in
the basement rock underneath the tunnel, and a subsequent strong seismic shaking from a
different largemagnitude seismic event that may occur years later. Non-linear finite elements
model the quasistatic fault rupture propagation through the soil overlying the bedrock and the
ensuing interaction of the rupture with the immersed tunnel. It is shown that despite an imposed
baserock offset of 2 meters, net tension or excessive compression between the tunnel
segments can be avoided with a suitable design of the joint gaskets. Then, the already
(permanently) deformed structure is subjected to strong asynchronous seismic shaking. The
tunnel is modeled as a 3D flexural beam connected to the soil through properlycalibrated
nonlinear interaction springs and dashpots, the supports of which are subjected to the freefield
acceleration time histories. The latter are basically obtained with 1D wave propagation
analysis, after they are modified to account for wave passage effects. The joints between the
tunnel segments are modeled with special non-linear hyper-elastic elements, while their
longitudinal pre-stressing due to the (7 bar) water pressure is also rigorously incorporated in the
analysis. The possibility of sliding is considered through the use of special gap elements. The
influence of segment length and joint properties is explored parametrically. A fascinating
conclusion emerges in all analysed cases for the joints between tunnel segments that were
differentially deformed due to static fault rupture displacements : upon subsequent strong
seismic shaking (with ground accelerations even exceeding 0.50 g), overstressed joints decompress and understressed joints re-compress, a healing process that leads to a more
uniform deformation profile along the tunnel. This is particularly beneficial for the precariously
de-compressed joint gaskets which thus avoid the danger of getting into net tension. Hence, the
safety of the immersed tunnel improves with subsequent strong seismic shaking !

INTRODUCTION

the narrowest crossing of the Corinthian Gulf


trench, which is characterized by an extensional
slip rate of about 1 cm/yr, the tunnel will
unavoidably cross at least one active fault. It
must be capable of safely undertaking the
ensuing permanent ground deformation.
Moreover, in the 100 year design life of the
tunnel, strong seismic shaking from another
earthquake, originating at a different major fault
located not directly under the site, is a distinct
possibility given the numerous active faults in
the Corinthian and Patraicos Gulfs, as well as in
their neighboring regions. Such seismic shaking

The objective of this paper is to investigate


the behaviour of a deep immersed tunnel under
the combined action of a major fault rupturing
underneath and a subsequent strong seismic
shaking. The presented research is part of a
feasibility study for a railway immersed tunnel in
the RionAntirrion straits, in Greece. The
construction of the tunnel will be a technological
challenge, due to the combination of great
water depth (67 m) (presently the world record),
very high seismicity, and deep alluvial soils. At

of 23 m x 11 m (width x height) section is


designed to accommodate two-way rail traffic.
The Rion-Antirron Straits is the narrowest
crossing of the Corinthian Gulf trench. The
trench is associated with a tensile tectonic
environment of NS direction, expressed by a
sequence of normal faults of EW strike dipping
towards the axis of the trench. Characterized by
a slip rate (extension) of almost 1 cm / yr, and
being one of the most seismically active zones
of Greece (Fig 1b), the area has experienced
five major earthquakes of Ms > 6, during the last
30 years. The most recent event was the Ms 6.2
Aegion 1995 earthquake, with an epicentral
distance of 28 km from the site. The Straits are
also affected by even more distant rupture
zones, such as the Ionian zone which is by far
the most active seismic zone in Europe,
capable of producing earthquakes of Ms > 7.
The recent Ms 6.4 Lefkada earthquake
(1482003) was only a small reminder of this
possibility.
Making use of the abundant geotechnical
data obtained in the course of design of the
neighboring RionAntirrion bridge we were able
to construct a credible soil profile of the
crossing. It consists of alternating layers of
sandy gravel to gravel, silty sand and clay, of
medium density and stiffness. The geotechnical
exploration reached a maximum depth of 100 m
below the seabed, without encountering
bedrock. Geological studies suggested that the
total thickness of the soil sediments exceeds
500 m. This was verified by a detailed
geophysical tomography conducted for the
needs of the present study [Tselentis, 2004]. It
was revealed that limestone bedrock lies at a
depth of about 800 m below seabed.

will find the immersed tunnel already injured


from the ground surface dislocation of the first
seismic event (which will have occurred
perhaps several tens of years earlier). It is
crucial for a successful design to ensure that
the permanent tensile deformation due to the
normalfault rupture, and the superimposed
subsequent dynamic deformation during
shaking will not jeopardize the water-tightness
of the tunnel at any moment. Figure 1 sketches
the two loading situations studied in the paper.
A FUTURE RIONANTIRRION RAILWAY LINK

A proposed railway link will be located at the


RionAntirrion straits connecting Central
Greece with Peloponnesus (Fig 2 and 3). The
straits will be crossed at approximately 300
meters east from the recently built cable-stayed
road bridge. At this narrowest point, the underwater length is about 2.5 km, with a maximum
depth of 67 m.
Two tunneling alternatives were initially
proposed and investigated : a continuous bored
tunnel, and a hybrid solution, combining a
central immersed tunnel at the deepest section
of the crossing, with two bored approach
tunnels at the two sides. The first solution would
require the boring of a tunnel at a maximum
depth of 95 m below the sea surface within
relatively loose soil. Facing water pressures of
the order of 10 bars relentlessly over a length of
at least 2 km was deemed far too risky, despite
the progress in shield tunneling technology, and
therefore only the second solution was chosen
for further research. The cross sections of the
immersed and the bored tunnel are depicted in
Figs 3b and 3c, respectively. The central tube

DYNAMIC (cyclic)
Decompression + Recompression

STATIC
Tensile Opening (Decompression)

accelerations i(t)

+
Fault Rupture

Figure 1: Definition of the problem : two types of loading that may take place in the life of the tunnel.

Lake Trihonis

Active Fault
Potentially Active Fault
0

10

20

(km)

Rion
Elik
i

Patras

Fau
lt

Aegion Fault

Figure 2: RionAntirrion straits with the active and potentially active faults

(a)

RION

ANTIRRION

Depth (m)

20

40

60

Embankment

Bored T
unn

Embankment
Immersed Tunnel

Special Connecting
Segment

80

l
Tunne
Bored

el

Special Connecting
Segment

4+000

4+500

5+000

5+500

6+000

Length (km + m)

1.5 m

0.5 m
5.5 m

11 m

4.0 m

4.4 m

4.4 m

1.5 m

2.0 m

1.0 m

1.5 m
3
1

0.7 m

4.4 m

4.4 m

2.5 m
4.0 m

23 m

12.5 m

(b)

(c)

Figure 3: (a) Longitudinal Section of the proposed tunnel, (b) Immersed tunnel cross-section, and (c) Bored
tunnel cross-section

ASEISMIC DESIGN OF THE PROPOSED RIONANTIRRION IMMERSED TUNNEL

As already discussed, the concrete


segments will be constructed in a dry-dock,
floated over the pre-excavated trench, and
lowered with the help of special sinking rigs.
Each segment will be lowered close to the
previous one, and brought to contact using
special guidance techniques. Once the two
segments gain contact, the water between them
will be drained, and the Gina gasket will be
compressed (Figs 4b and c) with the help of the
hydrostatic water pressure acting only at the
free side of the segment.
In our case, this pressure will be of the order
of 7 bars, and thus the use of the largest
available Gina gasket is indispensable. One
problem with the application of the method in
such depth is the extremely high compressive
load on the gasket. For the cross-sectional area

Until today, no immersed tunnel is known to


have ever been hit by an underneath rupturing
fault. On the other hand, two immersed tunnels
have been subjected to fairly strong seismic
shaking : the Bay Area Rapid Transit (BART)
tunnel in California, and the Osaka South Port
immersed tunnel in Japan. Both tunnels
behaved exceptionally well sustaining no
measurable damage.
After reviewing alternative seismic design
techniques already applied to existing
immersed tunnels, and given the good seismic
performance of the Osaka South Port immersed
tunnel, we proposed a similar design for our
immersed section (Fig 4).

BEFORE CONTACT

(b)

(a)

Gina Profile

Gina Profile
Tendon

Omega Profile

Shear Key

Tendon

Tendon

Coupler

(c)

Omega
Profile

Coupler

AFTER CONTACT

Figure 4: Aseismic design of the immersed section of the proposed railway link : (a) schematic detail of the
immersion joint, showing the gina gasket, the omega seal, the tendons along with the couplers, and
the shear key, (b) zoom-in of the immersion joint before the contact of two consecutive segments,
and (c) after the compression of the Gina gasket, installation of the Omega seal and connection of
the tendons with the use of a special coupler.

of our tunnel, 253 m2, the total hydrostatic force


to act on the gasket will be of the order of 175
MN ( 253 m2 x 700 kPa). Given the perimeter
(about 65 m) of the Gina gasket, the total force
per running meter will be 2.7 MPa. Gina
gaskets may fail in lateral tension under
significant
compressive
loads.
The
specifications of the largest available Gina
section show that they will be capable (in the
limit) of undertaking such a pressure. However,
since this will be the first Gina gasket to be
used in such depth, special analysis and testing
will be a prerequisite for acceptance.
After compression is completed, the
secondary omega seal is installed. Contrary to
the Gina profile that requires compression to
achieve water-tightness, the omega seal is
insensitive to the magnitude of compression.
However, none of the above materials can
transmit either shear or tension. To this end,
shear keys and tendons will be installed. A
shear key at the bottom of the tunnel will
transmit transverse shear forces, while two
similar shear keys at the side-walls will transmit
vertical shear. They are all constructed after the
segments are connected, by casting concrete in
situ. Tendons will transmit longitudinal tensile
forces, if necessary, acting as a secondary line
of defense.
The scope of this paper is to analyse the
behaviour of the tunnel under combined fault
induced displacement and strong seismic
shaking. The behaviour of the tunnel subjected
to different earthquake scenarios was
investigated in more detail in Anastasopoulos et
al [2007]. In this paper we parametrically
investigate three possible segment lengths: 70
m, 100 m, and 165 m. The segment length
determines the total number of the required
immersion joints, controlling their deformation
and the overall resilience of the tunnel. The
modern trend is to use longer segments, since
this leads to significant economy in construction.
For example, in the case of the recently built
3.5km resund Tunnel between Denmark and
Sweden [Mashall, 1999] the segment length
was chosen to be 175 m to minimize the cost. A
detailed study showed that lengths below 100
m and above 200 m could not be economical.
Naturally, this kind of optimization is very sitespecific and is also a function of the available
dry-dock facilities. In our case, the combination
of the 67 m depth with the vivid tectonic

environment do not really allow for this type of


optimization : our first priority would be to
secure its feasibility and devise a sound
aseismic design.
Besides the segment length (and hence the
total number of joints), the type of the Gina
gasket was also parametrically investigated.
The longitudinal deformation of the tunnel
depends mainly on the properties of the Gina
gasket. Additionally, since the Gina profile
constitutes the primary seal of the tunnel,
ensuring
its
permeability
is
critical.
Faultinduced
extension
and differential
settlement will unavoidably decompress at least
some of the joints. Additionally, during a strong
seismic event which may occur years later, the
longitudinal vibration of the tunnel will subject
the immersion joints to cyclic (dynamic) recompression
and
de-compression.
The
magnitude of the total de-compression is critical
for the design of the tendons. If the decompression is significant, the tendons will have
to undertake large tensile forces to ensure
impermeability.
Two types of joints (Fig 5) are parametrically
investigated. Type A refers to the idealized
behavior of the largest available Gina profile,
while type B is a proposed hypothetical doublesized Gina-type gasket. The behavior of this
hypothetical gasket provides wider deformation
limits, thus permitting significant additional
compression and decompression. The hyperelastic performance of both joints is estimated
from testing results of half-sized models
[Kiyomiya, 1995]. Already in 1995, Kiyomiya
had foreseen that in the future immersed
tunnels will be built deeper and in seismically
active areas : exactly this case. Our type B
gasket is the logical projection. Of course, it has
not yet been tested.
Finally, the allowance of the shear key is
parametrically investigated. It is expected that
this allowance will affect both the longitudinal
deformation of the tunnel due to fault-induced
deformation, and its transverse deformation
during the strong seismic shaking. If this
allowance is large enough, then the joints will
allow relative displacement and rotation
between two consecutive segments. On the
contrary, reduced allowance makes the
connection more fixed. Two extreme
alternatives are explored : 5 mm and 20 mm.

25 cm

20 cm

36 cm

Gina Type - A

Gina Type - B

40 cm

50 cm

Load (MN/m)

72 cm

10

20

30

40

Compressive Displacement (cm)


Figure 5: Hyperelastic behaviour of rubber gaskets used in the analysis. Type A refers to the largest
available GINA gasket, while Type B constitutes our hypothetical logical projection. The behaviour
has been estimated based on the half-scale tests of Kiyomiya [1995].

METHODOLOGY OF ANALYSIS

The analysis is conducted in three steps.


First, in Step 0 the hydrostatic pressure is
applied statically to the end of the segments,
to simulate the initial hydrostatic longitudinal
compression.
Then, in Step 1 the fault-induced
displacement profile is applied pseudostatically on the model, while
at the final stage, Step 2, the model is
subjected to the asynchronous dynamic
vibration. The methodology for the analysis
of fault rupture propagation, along with the
resulting seabed displacement profiles, is
presented in the next section. Then, in the
following
section,
the
asynchronous
dynamic analysis methodology is discussed
in detail.

The finite element code ABAQUS is utilized to


perform non-linear static and dynamic transient
analysis of the tunnel. The layout of the model
is depicted in Fig 6. The immersed tunnel is
modeled as a series of beams connected to the
soil through interaction springs and dashpots,
and with each other through an arrangement of
hyper-elastic elements simulating the behaviour
of the immersion joints. The tunnel segments
are modeled as beam elements. Special
transitional rigid elements are connected at
both ends of each tunnel segment to represent
the segment face, where the immersion joint is
placed.

Immersion Joint

Immersion Joint

ent
Segm m)
(bea

Fy

z
y

ay(x)

a x(x)

Immersion Joint

Fx

Immersion Joint

Fy
y

ks

Segment (beam)

Profile
Gina

Passive
Failure

slider

slider

Fy

c
Fx

Friction, x

k?

indu
Step 1 : Fault-

t
lacemen
ced disp

Shear Key
Allowance
xi-1 /C

xi /C

xi+1 /C

Step 2 : Asynchronous Seismic Shaking

Figure 6: Finite Element Modeling Layout : In the first step, the fault-induced displacements are applied to the
tunnel pseudo-statically. Then, in a second step we apply the asynchronous seismic excitation.

(expressed as force over displacement per unit


length of the tunnel [kN/m/m]), as functions of
the incident wavelength. For the vertical spring
stiffnesses St. John & Zahrah [1987] utilized the
fundamental solution to the problem of a
surfaceloaded halfspace rather than of a
within-loaded full-space Flamants problem,
instead of Kelvins (see Poulos & Davis, 1974 ;
Davis & Selvadurai, 1996).
The expressions of St. John & Zahrah
give the moduli of subgrade reaction (Winkler
spring stiffnesses) as functions of the
wavelength, a fact that introduces an additional
uncertainty. It is interesting to note that with the
ingenious introduction of wavelength they
accomplished the circumvention of the
singularity of a planestrain solution (both in
fullspace and on halfspace).
However, with immersed tunnels : (a) The
embedment is not sufficient for Kelvins solution
to even approximately apply; the half-space
solutions are more appropriate. (b) he

As foresaid, each tunnel segment is


connected to the overlying soil through
interaction springs and dashpots. The
acceleration time histories are applied to the
ends of these springs-dashpots with a time lag,
as described in detail in the relevant section.
SoilTunnel Interaction Parameters
Assigning proper values to the horizontal (x, y)
and vertical (z) supportingspring constants
(Winkler moduli) is a task with substantial
uncertainty. For embedded tunnels, St. John &
Zahrah [1987] derived an expression based on
integration of Kelvins fundamental solution for
a concentrated (point) load acting within an
infinite elastic medium (full-space), assumed to
be homogeneous and isotropic. Upon
computing the settlement due to load
distributed uniformly across the width of the
tunnel and sinusoidally over a wavelength along
its axis, they derived approximate expressions
for the two horizontal spring stiffnesses

relatively high rigidity of each tunnel segment


(of aspect ratio L/B 3 to 7 in our study) with
respect to the nearsurface (usually soft) soil
will lead to deformation of the segment in
horizontal translation (in the x or y direction)
that may be closer to uniform than to sinusoidal;
hence, the rigid-foundation solution is more
appropriate. (c) The singularity of the planestrain solution exists only in the static problem,
not the dynamic. One additional factor
suppresses the singularity in this case : the
soil modulus is not constant, but increases with
depth [Gazetas, 1983].
Therefore, one can justifiably utilize the
wealth of published elastodynamic solutions for
a rigid long rectangular foundation on halfspace, to obtain not only Winklertype springs
but also dashpots to represent the elastic soil
structure interaction. In this case the
approximate
expressions
for
a
nonhomogeneous half-space proposed by
Gazetas (1991) are utilized. To this end, using
as a starting point the generic profiles of Fig. 4,
we eventually model the effective shear
modulus profile as :

These expressions are considered valid for all


frequencies a reasonable simplification for
translational modes of vibration (e.g. Gazetas,
1991). Moreover, they ignore the effect of
embedment ; this effect is added indirectly
only in the lateral (y) direction (Fig. 6), by
increasing the spring coefficient, ky , by the
elastic sidewall resistance:

ky Efill h/B

where Efill = the (average) Youngs modulus of


the backfill, and h = the effective depth of
embedment (assumed to equal 2/3 of the actual
maximum depth, 4.4 m, of Fig. 2).
A similar approach was advocated by
Vrettos [2005], who also expressed the spring
stiffnesses as constants, independent of
wavelength or frequency.
The vertical, cz , lateral, cy , and
longitudinal, cx , dashpot coefficients, reflecting
the radiation and hysteretic damping in the soil,
are similarly obtained from the expressions and
diagrams of Gazetas [1991]. In view of the
strong soil inhomogeneity and the relatively low
dimensionless frequency parameters, ao = (2
/T) B / Vso , that are of prime interest here,
these coefficients play a minor role in the
response and are not further discussed here.
The tangential contact forces transmitted
from the seabed to the tunnel are limited by
friction
and
passive
resistance.
As
schematically illustrated in Fig. 6, in the
longitudinal direction (x) the behavior of the
interface is approximated with that of a simple
slider of a friction coefficient x. In the
transverse direction (y), the interface is more
complex, with sliding accompanied by passive
type deformation of the backfill. Therefore, the
equivalent friction coefficient, y , is estimated
through 2D plane strain analysis of the tunnel
cross-section.
The (immersion) joints between the tunnel
segments are modeled with non-linear springs.
In the longitudinal direction (x), the springs refer
to the Gina gasket. As already discussed, their
hyper-elastic
restoring
forcedeformation
backbone curves (Fig. 5) are consistent with the
results of half-size model tests [Kiyomiya,
1995]. In the transverse (y) and vertical (z)
directions Gina-type gaskets cannot transfer
shear ; the drift of the tunnel depends solely on
shear key allowance. Thus, the behavior of the

z
G( z ) Go 1 +

(1)

where Go is the shear modulus at z = 0 ; B =


23.5 m is the width of the tunnel ; and , n soilmodel parameters. The three parameters, Go ,
, and n , were obtained by curve fitting not the
initial Gmax profile, but the equivalentlinear G =
profile, i.e. the profile of the shear
modulus in the last iteration of the equivalent
linear wave propagation analysis. The
distributed vertical and horizontal springs are
then obtained as the static vertical, lateral, and
axial stiffnesses of a very long tunnel.
Expressed as stiffnesses per unit length, the
Winkler moduli kz , ky , kx , (in kN/m/m) in terms
of Poissons ratio ( 0.50) are:

kz

0.73
n
Go (1 + 2 )
1 v

(2)
n

2
2
ky
Go 1 +
2 v 3

0.2
kx k y
0.75 v

(3)
n

B 1
1 Go 1 +
L 2

(5)

(4)

1994].
Based
on
the
aforementioned
geophysical investigation, the central part of the
straits, down to the depth of 800 m where the
bedrock was detected, is discretized. The
analysis is conducted in plane strain, with the
use of the FE code ABAQUS. The model is
displayed in Fig 7b, with an H = 800 m thick soil
layer, at the base of which a normal fault
dipping at an angle ruptures and produces a
downward movement of vertical amplitude h.
he total width of the model is B = 4H = 3200
m, following the recommendation of Bray [1990]
that a B : H = 4 : 1 ratio is sufficient to minimize
undesirable boundary effects. At the central
1600 m of the model, the discretisation is finer,
with the quadrilateral elements being 20 m x 20
m (width x height), while a suitable slip-line
tracing algorithm reduces the element size in
the neighborhood of the rupture path. At the two
edges of the model, where the deformation is
expected to be much smaller, the mesh is
coarser : 40 m x 20 m. The differential
displacement is applied to the left part of the
model in small consecutive steps.
Several experimental and numerical studies
have shown that soil behaviour after reaching
failure is a decisive factor in rupture
propagation.
After a thorough review of the literature, an
elastoplastic constitutive model was adopted
[Anastasopoulos & Gazetas, 2007]: MohrCoulomb failure criterion, with an isotropic strain
softening rule, applied to cohesion c , angle of
friction , and angle of dilation . Denoting f
the plastic shear strain at which soil reaches its
residual strength, we consider c, and as
linearly decreasing functions of the total plastic
strain until they reach their residual values cres,
res, and res. Typical values of f range from
5% to 15%. Equally important is the yield
strain y, which depends on both strength and
shear stiffness.
A detailed study on fault rupture propagation
conducted by Anastasopoulos & Gazetas
[2007] has successfully compared the
numerical
results
with
case-histories,
experimental data, and earlier numerical studies.
Additionally, a successful Class A prediction
was published on the Internet utilizing
centrifuge
experiments
performed
on
Fontainebleau sand at the University of Dundee,
as part of a European research project [Davies
& Bransby, 2004]. This, along with additional

joint is modeled with special gap elements,


which would only transmit shear after the shearkey allowance closes, in which case their
stiffness becomes very large, depending mainly
on the stiffness of the concrete section in the
area of the shear key (Fig. 6, bottom left).
ANALYSIS OF FAULT RUPTURE PROPAGATION

We first analyse the fault rupture


propagation through the overlying soil deposit.
As it was previously mentioned, the location
and the magnitude of the potential dislocation of
the ground surface depend not only on the type
and magnitude of the fault rupture, but also on
the geometry and material characteristics of the
overlying soils. Geotechnical exploration in
Rion-Antirrion only reached a maximum depth
of 100 m below the seabed, without
encountering bedrock. To shed more light in the
tectonic environment of the area, and to gain
better knowledge of the alluvium thickness, a
detailed
geophysical
investigation
was
conducted as part of our feasibility study
[Tselentis, 2004]. Before proceeding with the
analysis of fault rupture propagation it is
necessary to discuss briefly the most important
findings.
Given the high seismicity of the area, the
passive tomography method was applied. A
dense network of 70 seismographs was
installed in the project area, to capture seismic
waves oriGinating from micro-tremors (small
magnitude
earthquakes).
A
tomography
imaging technique was used to locate
potentially active faults. The seismographic
network indicated no seismic activity at depths
less than 1.5 km, verifying the hypothesis that
faults within the straits do not outcrop on the
seabed. As depicted in Fig 7a, five rupture
zones were detected within the straits, with
direction practically perpendicular to the tunnel
axis. Close to the side of Antirrion, the mapped
Antirrion fault was also detected. With an
extension slip rate of 0.7 mm/yr, it constitutes
the most important tectonic feature in the
immediate neighborhood.
The finite element method has proven
successful
in
analysing
fault
rupture
propagation through soil provided that certain
conditions are satisfied, such as (i) the use of a
very refined mesh in the neighborhood of the
potential rupture, and (ii) the use of a suitable
nonlinear constitutive law for the soil [Bray,

In both cases the shear modulus G was


assumed linearly increasing with depth while y
is kept constant. The dense soil reaches
failure at relatively low strains (brittle
behaviour), while the loose soil accommodates
relatively high strains before yielding (ductile
behaviour).

verification of the developed models against


case histories and experiments from the
published literature, provided confidence in
using our numerical methodology.
Conservatively assuming a dip angle of 45,
a downward displacement of vertical magnitude
h is imposed on the left half of the model (the
hanging wall). As already mentioned, the soil
profile is known only for the upper 100 m :
alternating layers of sandy gravel to gravel, silty
sand to sand, and silty clay to clay. Given these
uncertainties, we parametrically investigated
two idealized soils :
(a) dense cohesionless soil : = 45, res =
30, = 15, res = 0, and y = 1.5%
(b) loose cohesionless soil : = 30, res = 25,
= 5, res = 0, and y= 5.0 %

To be on the conservative side, despite the


optimistic geophysical results, we assumed h =
2 m as the design bedrock displacement
(vertical component). Note that for very small
values of the base fault displacement h, relative
to the soil thickness H, the rupture cannot reach
the ground surface. Given the significant 800 m
depth, a bedrock displacement of 2 m
corresponds to an h/H = 0.25 %. The required

Distance (km)
-4
0

-2
2000

Immersed Tunnel

3000

1000

5000

(a)

Vp (m/s)

Fa u
lt

2000

on

Depth (m)

4000

4000

An
tirr
i

3000

RION

ANTIRRION

B = 4H = 3200 m

H = 800 m

Hanging wall

(b)

Figure 7: (a) Mapping of active faults in the area of the crossing, based on the geophyical tomography
[Teslentis, 2004], (b) Finite element discretization for the plane strain analysis of fault rupture
propagation through the 800 m soil sediment.

10

overlying structures. The horizontal strain, x ,


along the ground surface (positive values are
for tension) is depicted in Fig 9d.
Observe that the ground surface deforms
smoothly, as the rupture cannot reach the
surface. The bending deformation, expressed
through the distortion angle (Fig 9c), reaches
its peak, 0.26%, at a distance d 200 m from
the straightline projection of the rupture. On the
other hand, the horizontal (tensile) deformation
is maximized, x = 0.25%, at a distance d 500
m (Fig 9d). The computed herein displacement
profiles (x and y) are used as the input
displacement to be imposed on the tunnel
model. The tensile deformation causes
decompression of the joints, while the bending
deformation decompresses some of the
immersion joints and possibly compresses
further some other. Since the maximum
bending and tensile deformation does not occur
at the same location, we identify two tunnel
fault rupture relative positions (Fig 9) :
(1) Position 1 : the center of the tunnel
coincides with the location of max ,
and
(2) Position 2 : the center of the tunnel
coincides with the location of max x .
The prediction of the worst-case scenario is
not straightforward. Therefore, both positions
are investigated.
Typical results of a complete analysis are
portrayed in Figure 10, in terms of joint
deformation, sliding displacement, bending
moments and axial forces. Figure 10a
corresponds to segment length L = 70 m with a

h/H ratio for the outcropping of the rupture is an


increasing function of soil ductility, expressed in
our modified MohrCoulomb constitutive law,
through y. Even for a the brittle soil (a) with y
= 1.5 %, an h/H in the order of 0.75% is
required for the fault to break out
[Anastasopoulos & Gazetas, 2007].
EFFECTS OF FAULT RUPTURE ON THE TUNNEL

Characteristic pictures of the propagating


rupture in a deposit of idealized dense sandy
soil (y = 1.5 %) is given in Fig 8 in the form of
four snapshots of the deformed mesh, along
with the accumulated plastic strain, for different
bedrock displacements. Darker regions denote
higher plastic strains. Observe that for bedrock
displacement up to h = 4 m (i.e. h/H = 0.50%),
the propagating rupture does not emerge on the
surface (seabed). Only when the imposed
bedrock displacement exceeds 6 m (i.e, h/H =
0.75 %) does the dislocation (barely) emerge
on the surface. For the design displacement of
2 m the rupture is clearly far from reaching the
seabed (model surface). The situation is even
better with the idealized loose sandy soil : an
h/H of the order of 1.5% is needed for the fault
to outcrop [Anastasopoulos, 2005].
Figs 9a and 9b portray the profiles of vertical,
y, and horizontal displacement, x, along the
ground surface, for the (conservative) dense
soil and for hd = 2 m. The horizontal distance, d,
is measured from the point of application of the
bedrock displacement. Fig 9c depicts the slope
of the ground surface a useful response
parameter in assessing the damage potential to

h=1m

h=2m

h=4m

h=6m

Figure 8: Snapshots of deformed mesh and plastic strain for a fault rupture = 45o in dip and bedrock
displacements h/H = 1, 2, 4, and 6 m, for the idealized dense sandy soil.

11

Horizontal distance d (m)


-400

400

800

-1600 -1200 -800


0

1200 1600

-0.5

-0.5

-1

-1

x (m)

y (m)

-1600 -1200 -800


0

Horizontal distance d (m)

-1.5

-400

800

1200 1600

-1.5

-2

-2

Position 2

Position 1
-2.5

-2.5
0.4

0.4

Position 1

0.3

0.3

0.2

0.2

x (%)
( )

(%)

400

0.1

Position 2

0.1
0

-0.1
-1600 -1200 -800

-400

400

800

1200

1600

Horizontal distance d (m)

-0.1
-1600 -1200 -800

-400

400

800

1200

1600

Horizontal distance d (m)

Figure 9: Fault rupture propagation analysis results for a fault rupture = 45o in dip and bedrock
displacements hd = 2 m, for the idealized dense sandy soil : (a) horizontal displacement x, (b)
vertical displacement y, (c) angular distortion , and (d) horizontal strain x at the seabed.

2. This initial slippage is greater in the middle


of the tunnel, where the maximum tensile
deformation x takes place. Increasing the
segment length only marGinally increases
the maximum sliding displacement (from 7.5
cm to 9 cm in the middle of the tunnel for
segment lengths L = 70 m and L = 100 m
respectively).
3. At first, the axial force exhibits an initial prestressing of 160 MN due to hydrostatic
compression (Step 0). Then the application
of the faultinduced deformation (Step 1)
causes longitudinal decompression of the
tunnel, reducing the axial force (N)
significantly. In the case of the 70 m
segments, N drops from 160 MN to only 30
MN near the middle. The increased
segment length makes things worse: the
initial compression is completely lost, and
even some (small) tensile stressing
develops (at point C).

type A gasket, while Figure 10b to segment


length L = 100 m with the type B gasket.
Results for L = 165 m have been also examined
but are not shown herein: with such a large
length the tunnel cannot sustain the total
developing stressing without several joints
experiencing net tension a precarious
situation indeed. Similarly, the allowance in
the shear keys plays only a minor role; thus,
results are given only for a single value (5 mm)
of this allowance. In all cases the results
presented herein correspond to the position 2
scenario since this proved to be the most critical
case.
The following conclusions are drawn:
1. The application of the faultinduced tensile
displacement decompresses all the joints,
with those near the middle of the tunnel
experiencing the greatest tension with more
pronounced being the decompression when
the segment length is 100 m (Figure 6b).

12

methods: (i) making use of the widely accepted


equivalent linear approximation code SHAKE
[Schnabel et al, 1975], and (ii) applying the
nonlinear
constitutive
model,
BWGG,
developed by Gerolymos & Gazetas [2006].
Having estimated the free-field acceleration
time histories at the seabed, we apply the
methodology of EC8 [2002] to impose the
acceleration time histories on the tunnel
supports with a time lag. Given the length of the
tunnel ( 1 km), seismic shear waves incident at
an angle different than 90o to the horizontal,
appear as traveling with a finite propagation
velocity, C , along the surface (and hence
along the tunnel). This time lag leads to
asynchronous vibration of the tunnel segments.
It is computed through the apparent wave
velocity C. If we denote with x the longitudinal
axis of the tunnel, at a distance xi from the

4. The
fault-induced
deformation
is
responsible for the development of
longitudinal bending moments, My. With L =
70 m, at point A a moment of 70 MNm is
developed while point C (at the opposite
end) is almost unaffected by the imposed
displacement (My = 5 MNm). The increase
of L to 100 m leads to an increase of the
longitudinal bending moment.
METHODOLOGY OF DYNAMIC ANALYSIS

Fig 11 illustrates schematically the methodology


used in our dynamic analysis. The immersed
tunnel is modeled as a beam connected to the
soil through interaction springs and dashpots.
Free-field acceleration time histories are
computed trough 1-D wave propagation
analyses, that were conducted with two

(a)

CC

BB

70 m segment length- type B


A gasket
Joint
Decompression (cm)

Segment Slippage
(cm)

N (MN)

M (MNm)

-130

-70

16

7.5

-30

40

10

-50

(b)
AA

C C

BB

100 m segment length- type B gasket


Joint
Decompression (cm)

Segment Slippage
(cm)

N (MN)

M (MNm)

-100

-150

21

-9

-80

-10

14

-2

10

140

Figure 10: Joint deformation, sliding displacement, axial forces and bending moments after the fault induced
displacements for (a) 70 m segments and (b) 100 m segments. Fault at Position 2, type B gasket,
and 5 mm shear-key allowance.

13

which compares quite well with the conservative


design spectrum of the neighboring Rion
Antirrion cable-stayed bridge. The (oriGinal)
Aegion 1995 and the Lefkada 2003
accelerograms are among the strongest ever
recorded in Greece. The Aegion accelerogram,
at a distance of only 28 km from the site, is
characterized by a single long-period pulse of
0.54 g one of the greatest longperiod PGAs
ever recorded in Greece. The Lefkada
accelerogram (recorded at a distance of 120 km
from the straits), is characterized by a PGA of
0.43 g and six strong cycles with acceleration in
excess of 0.30 g.
Two types of 1-D verticalwave propagation
analyses were conducted : (a) by applying the
widely accepted equivalent linear method of
analysis with the SHAKE code [Schnabel et al,
1975], and (b) by applying a nonlinear inelastic
constitutive model for the soil, coded in the
program NLDYAS [Gerolymos & Gazetas,

beginning, the seismic excitation will arrive with


a time lag ti = xi / C (Fig 11). The apparent
wave velocity was conservatively assumed
equal to 1000 m/s (Eurocode EC8 Part 2
(Bridges) suggests apparent wave velocities C
> 1500 m/s, even for the softest soil category).

Analysis of Soil Response

As depicted in Fig 12, three real earthquake


records were used for the 1-D wave
propagation analyses : (a) the JMA record in
the 1995 Ms 7.2 Kobe earthquake, (b) the
(unique) record of the 1995 Aegion Ms 6.2
earthquake, and (c) the (unique) record of the
2003 Lefkada Ms 6.4 earthquake. They were all
scaled to a peak acceleration of 0.24 g, as
specified by the Greek Seismic Code [EAK
2000] for the area of study. The first record,
Kobe JMA, has a (scaled) response spectrum
L = 980 m

x
a

xi

t
EC8

xj

xi /C

xj /C

L /C

C : apparent wave velocity


at distance xi , time lag ti = xi/C

0.48 g

SHAKE

G
G

Depth (m)

100

WGG
Dense
Sand

High Plasticity
Clay

Loose
Sand
0

0.24 g

250

500

750

1000

1250

Vs (m/sec)

Kobe JMA

Figure 11: Dynamic Analysis methodology : The bedrock acceleration is analysed in 1-D to derive the
acceleration at the seabed. Then, following the EC8 methodology, the acceleration time histories
are imposed with a time lag on the tunnel model.

14

while the Lefkada accelerogram yields 0.63 g.


Five more possible soil profiles were
parametrically investigated. From all of the
examined soilprofile and seismicexcitation
scenarios, the PGAs were found to range
between 0.40 g and 0.63 g. The three
computed seabed accelerograms of Fig 12
were used as the free-field input for the
dynamic analysis of the tunnel. Interestingly,
the design acceleration levels of the Rion
Antirrion (cablestayed) Bridge (PGA 0.50 g)
are similar to the PGA values of this figure.

2005]. The equivalent linear is definitely the


most popular method, having however certain
limitations, especially in the case of very deep
deposits (as in our case). On the other hand,
most nonlinear soil models are incapable of
simultaneously reproducing the observed shear
modulus decrease and damping ratio increase,
usually overestimating the hysteretic damping
at large strains (if the Masing rule for
unloadingreloading is used). The recently
developed NLDYAS code uses a constitutive
model, denoted BWGG, which avoids such
disadvantages, and is capable of reproducing
even some of the most complex non-linear
characteristics of cyclic behavior, such as cyclic
mobility and liquefaction.
Fig 12 illustrates the 1-D wave propagation
results for one of the worsecase scenarios :
stiff soil profile with Vs = 1300 m/s below 100
m depth. The Aegion and the Kobe JMA
records produce peak ground acceleration
(PGA) in the order of 0.50 g at the seabed,

COMBINED EFFECTS OF FAULT RUPTURE AND


(SUBSEQUENT) DYNAMIC EXCITATION

Typical results of a complete analysis are


portrayed in Figs 1316, in the form of time
histories of accelerations, internal forces,
displacements
and
deformations,
which
develop at various points along the tunnel
during the strong seismic shaking . Static

0.6

0.6

0.3

0.3

-0.3

-0.3

-0.3

a (g)

0.48 g

-0.6
0

0.3

-0.6

10

-0.6

15

0.63 g

0.6

0.50 g

t (sec)

10

15

t (sec)

250

10

15

t (sec)

Profile C (hard)

100

1300

250

500

750

1000 1250

Vs (m/sec)

0.6

a (g)

JMA Kobe 1995

0.6

Aegion 1995

0.24 g

0.6

0.3

-0.3

-0.3

-0.3

0.24 g

-0.6

-0.6

-0.6

10

t (sec)

15

Lefkada 2003
0.24 g

0.3

0.3

10

t (sec)

15

10

15

t (sec)

Figure 12: 1-D dynamic wave propagation analysis results for the worst-case scenario : stiff soil profile,
with Vs = 1300 m/s below 100 m depth.

15

does not appear to have a significant effect on


the longitudinal accelerations of the tunnel.
In the transverse direction the response is
also differentiated along the length of the tunnel.
At the two ends (B, D), the acceleration time
histories are quite similar to the input motion (A).
At these points, the immersed tunnel is forced
to follow the input excitation, as it is rigidly
connected to the bored approach tunnels. The
latter are as a first approximation assumed to
follow the imposed displacements. However,
the central part of the tunnel (C) exhibits a
different behaviour. With the exception of a few
extremely high-frequency acceleration spikes,
attributable to gap closures at the shear-keys,
the acceleration is cut-off at about 0.30 g, a
value equal to the transversal friction
coefficient y. Hence, this acceleration cut-off
implies transverse sliding of the tunnel. As
expected, this sliding takes place only near its
center, as the two ends are restrained by the
bored tunnels. Again, the segment length does
not affect significantly the lateral acceleration of
the tunnel. It is also noted (although not shown
here) that the tunnel response accelerations are
insensitive to the gasket type.
The sliding of the tunnel (longitudinally and
transversally) is evident in Fig 14. In the
longitudinal direction, the tunnel segments have
already slipped before the application of the
dynamic excitation (Step 2), due to the fault
induced longitudinal deformation (Step 1). This
initial slippage is greater in the middle of the
tunnel, where the maximum tensile deformation
x takes place, and increases with segment
length (from 7 cm in the case of the 70 m
segments, to about 12 cm for the longer
segments). The increase of segment length
also leads to increased additional dynamic
relative displacements. While for the 70 m
segments the additional sliding due to the
dynamic excitation reaches 10 cm, in the case
of the 100 m segments the maximum relative
dynamic displacement exceeds 17 cm (location
B). In all cases, however, the permanent
(residual) displacement is significantly lower
than the maximum value. In the transverse
direction, the central segments of the tunnel
experience substantial sliding, while the ends
keep pace with the overlying soil just as
was the case with the transversal acceleration
time histories in Fig 13. Increasing the segment
length (from 70 m to 100 m) only marginally

permanent deformation and distress is


assumed to have already taken place in a
preceding event of fault rupturing underneath
the site of the tunnel.
Each of these figures presents results for
two different values of segment length, 70 m
and 100 m . Results for 165 m, which was also
examined, are not shown : with such a large
length the tunnel cannot sustain the total
developing stressing without several joints
experiencing a net tension a precarious
situation indeed.
Only results for the (scaleddown) Kobe
JMA excitation are shown in the said figures.
Although using the other motions as excitation
leads to slightly different results, the qualitative
behaviour of the tunnel remains the same and
the conclusions drawn herein do not change.
Similarly, the allowance in the shear keys
proved to play only a minor role ; thus, results
are given only for a single value (5 mm) of this
allowance . Finally, only one location of the
emerging fault rupture, position 2, is examined
in the aforementioned figures. This is the most
critical of the two studied locations.
Each figure presents the results for the
longitudinal (x) and the transverse (y)
response ; the former shown above and the
latter below the sketch of the tunnel.
Specifically, the details of each figure are as
follows :
Fig 13 shows the time histories of
acceleration which develop in the middle of
three segments of the tunnel (B,C,D), and
contrasts them with the ground excitation (point
A). Observe that in the longitudinal direction (x)
the response of the tunnel differs from point to
point. The peak acceleration is amplified from
0.48 g at the base to approximately 0.60 g near
the terminal segments (at the two edges of the
tunnel, B and D). Near the center, however, the
longitudinal acceleration is cut-off to about 0.30
g, implying sliding. This differential longitudinal
response is attributed to the previous non
uniform (along the axis) displacement from the
fault rupture and the ensuing tendency of the
(springrestrained)
tunnel
segments
to
readjust under the dynamic loading. This is
explained in detail in the next section. If the
faultinduced displacement did not pre-exist
(Step 1), then the dynamic response of the
tunnel, in the longitudinal direction, would be
nearly uniform. The increased segment length

16

0.60

0.6 A (base)

0.6

0.3

0.3

0.3

0.3

ax (g) 0

-0.3

-0.3

-0.3

-0.3

0.48 g

-0.6

0.6

-0.6

-0.6

10

0.6

-0.6

10

10

10

10

10

t (sec)
B
(a)

A
0.6

0.6

A (base)

0.6 C

0.6

0.3

0.3

0.3

0.3

ay (g) 0

-0.3

-0.3

-0.3

-0.3

-0.6

-0.6
5

10

-0.6
5

10

-0.6
5

10

t (sec)
0.6

ax (g)

0.6

A (base)

0.60

0.6

0.6 D

0.3

0.3

0.3

0.3

-0.3

-0.3

-0.3

-0.3

0.48 g

-0.6
0

-0.6

-0.6
0

10

10

-0.6
0

10

t (sec)

(b)
A
0.6

ay (g)

0.6

A (base)

0.6

0.6

0.3

0.3

0.3

0.3

-0.3

-0.3

-0.3

-0.3

-0.6

-0.6
0

10

-0.6
0

10

-0.6
0

10

10

t (sec)

Figure 13: Longitudinal and transverse acceleration time histories for : (a) 70 m segments, and (b) 100 m
segments (Kobe JMA excitation, fault at Position 2, type B gasket, and 5 mm shear-key
allowance).

17

x (cm)

30 A

30 B

20

20

10
0

30
20

20

10

10

10

17 cm

-10

-10
0

30 D

-10
0

10

-10

10

10

10

t (sec)

(a)

y (cm)

10

10 B

10

10

-5

-5
0

10

-5
0

10

-5
0

10

10

10

x (cm)

t (sec)
27 cm

30 A

30

20

20

20

20

10

10

10

10

30

-10

-10
0

-10
0

10

30 D

10

-10
0

10

t (sec)

(b)

y (cm)

10 A

10 B

10 C

10 D

-5

-5
0

10

-5
0

10

-5
0

10

10

t (sec)

Figure 14: Longitudinal and transverse sliding displacements for : (a) 70 m segments, and (b) 100 m
segments (Kobe JMA excitation, fault at Position 2, Type B gasket, and 5 mm shear-key
allowance).

18

thick cross-section of the tunnel can easily


undertake such a stressing, even with the
minimum allowable reinforcement. While My is
insensitive to the thickness of the rubber gasket
(Type A versus B), the increase of joint
allowance to 2 cm relieves the segments from
roughly 10 % of the transverse bending.
The longitudinal and transversal deformation
(x and y) of the immersion joints are the most
crucial response parameters for the seismic
safety of an immersed tunnel. Their time
histories during the seismic shaking after the
fault rupture has already had its effect are
portrayed in Fig 16, for the 70 msegment
tunnel. We compare the response, of the joints
equipped with the Type A gasket, versus that of
the joints with Type B gasket. Notice that in the
longitudinal direction, the Gina gasket
experiences an initial hydrostatic compression
(Step 0) of 17 cm and 28 cm, for Type A and B
gasket, respectively. The previous application
of the faultinduced tensile displacement
opens-up all the joints, with those near the
middle of the tunnel experiencing the greatest
tension. In the case of the Type A gasket, the
17 cm hydrostatic compression is practically
completely lost at the central part of the tunnel
(point C), where the leftover compression after
the fault rupture amounts to merely 2 cm. The
situation near the terminal segments A and D is
less critical tragic, with the remaining
compression of the gasket being 14 cm and 8
cm, respectively. During shaking, the gaskets
experience
alternating
cycles
of
decompression and re-compression. While the
dynamic compression is acceptably small, there
is some tensile deformation in gaskets A and C
a rather precarious situation.
The use of Type B rubber gasket improves
substantially the performance. One of the
reasons, of course, is the 28 cm of initial
hydrostatic compression. Much greater margins
for de-compression are thus available. The
fault-ruptureinduced de-compression of the
joints, still leaves a minimum of 12 cm
compression at the critical central part of the
tunnel,
C.
As
expected,
the
larger
decompression is observed at joints near the
middle of the tunnel, where the fault-induced
tensile deformation of the seabed is maximum ;
near the end segments unloading of gaskets is
substantially less. During dynamic oscillation,
the rubber gaskets experience cycles of

increases the maximum sliding displacement


(from 9 cm to 11 cm). In all cases, the
permanent (residual) sliding displacement does
not exceed 3 cm near the center of the tunnel,
and it is practically negligible at the two ends.
The transverse sliding displacement is also
insensitive to the thickness of the rubber gasket
(Type A versus B).
Fig 15 illustrates the bending moments and
axial forces in characteristic cross-sections of
the immersed tunnel. At first, the axial force
exhibits an initial pre-stressing of 160 MN due
to hydrostatic compression (Step 0). Then the
application of the faultinduced deformation
(Step 1) causes longitudinal decompression of
the tunnel, reducing the axial force (N)
significantly. In the case of the 70 m segments,
N drops from 160 MN to only 30 MN near the
middle. The increased segment length makes
things worse : the initial compression is
completely lost, and even some (small) tensile
stressing develops (at point C). The
readjustment of the segments during dynamic
shaking (thoroughly explained in the sequel),
leads to a redistribution of the longitudinal axial
force. Segments with higher axial forces tend to
get unloaded, while those subjected to tension
(smaller compression) tend to attract higher
compression. At the end of shaking, this
redistribution leads to a nearly uniform
distribution of the axial force : about 60 MN in
the case of the 70 m segments and 40 MN for
the 100 m segments. During shaking, a
maximum net tension of 10 MN takes place with
the 100 m segments (A).
Apart from the axial forces, the fault-induced
deformation is also responsible for the
development of longitudinal bending moments,
My . With the 70 m segments the application of
the faultinduced deformation creates a
bending moment of 60 MNm, which during
seismic shaking fluctuates by barely 15 MNm.
Increasing the segment length to 100 m leads
to a significant increase of My to 150 MNm, with
no effect on the dynamic fluctuation. The
bending moments are also insensitive to the
type of the rubber gasket.
The transverse bending moments, Mx are
not plotted here. Such moments do not develop
from the (normal to the axis) faultrupture.
Increasing the segment length from 70 m to 100
m, significantly increases My from 250 MNm to
400 MNm. In all cases, the 23 m x 11 m, 1.5 m

19

tt (sec)
(sec)
-2

N (MN)

-2

10

10

-2
0

-50

-50

-50

-100

-100

-100

-150

-150

-150

-200

F*

Dynamic

F*

-200

Dynamic

10

F*

-200

Dynamic

(a)

My (MNm)

200

200

200

100

100

100

-100

-100

F*

-200

-2

Dynamic
0

-100

F*

-200
8

10

-2

Dynamic
0

F*

-200

Dynamic

10

-2

10

10

-2

10

10

t (sec)

* : fault induced displacement

t (sec)
-2

N (MN)

-2

10
0

-50

-50

-50

-100

-100

-100

-150

-150

-150

F*

-200

Dynamic

F*

-200

Dynamic

F*

-200

Dynamic

(b)

My (MNm)

200

200

Dynamic

200

100

100

100

-100
-200

-100

F*
-2

-100

F*

-200
0

10

-2

Dynamic
0

F*

-200
8

10

-2

Dynamic
0

t (sec)

Figure 15: Axial Force and Bending Moment time histories for : (a) 70 m segments, and (b) 100 m segments
(Kobe JMA excitation, fault at Position 2, type B gasket, and 5 mm shear-key allowance).

20

x (cm)

30 A

30

30

20

20

20

20

10

10

10

10

0
0

30

0
0

10

10

0
0

10

10

t (sec)

(a)

y (cm)

10

10 B

10

10

-5

-5

-5

-5

-10

-10
0

10

-10
0

10

-10
0

10

10

x (cm)

t (sec)
30 A

30 B

30 C

30 D

20

20

20

20

10

10

10

10

0
0

0
0

10

10

0
0

10

10

10

t (sec)

(b)

y (cm)

10 A

10 B

10 C

10 D

-5

-5

-5

-5

-10

-10
0

10

-10
0

10

-10
0

10

t (sec)

Figure 16: Longitudinal and transverse joint deformation for : (a) 70 m segments with Type A gasket, and
(b) 70 m segments with Type B gasket (Kobe JMA excitation, fault at Position 2, and 5 mm
shear-key allowance).

21

coinciding with the location of maximum x), the


tectonic deformation leads to a de-compression
of all the joints (Figs 17a, c, and d). On the
other hand, when the tunnel is placed at
Position 1 relative to the fault rupture (center of
the tunnel coinciding with the location of max.
), most joints de-compress, except those near
the left edge of the tunnel (Fig 16b). At this
location, the sagging deformation of the seabed
seems to be the most important, imposing some
further compression of the joints. More
specifically, in the case of the 70 m segment
tunnel with the ( thick )Type B rubber gaskets
at Position 2 (Fig 16a), the joints decompress
from their initial hydrostatic compression of 28
cm to a minimum of 12 cm (a decompression of
16 cm) near the center of the tunnel (where
maximum tensile deformation occurs). This
decompression is reduced significantly near the
left edge of the tunnel (x 26 cm), but not to
the same extent at the right (x 18 cm).
As seen in Fig 17c, the thickness of the
rubber gasket does not substantially affect the
de-compressed
profile.
But
while
decompression is practically the same, in the case
of the (slimmer) Type A rubber gasket, the initial
hydrostatic compression is only 18 cm. As a
result, the joints near the center almost loose all
their pre-stressing (x 2 cm). This is
threatening for the watertightness of the tunnel !
Increasing the segment length to 100 m
(combined with a Type B gasket) improves the
safety margin : xmin 6 cm near the center of
the tunnel (Fig 16d) : This effect can be easily
explained : while in the case of the 70 m
segments the imposed deformation is absorbed
by a total of 13 joints, in the case of the 100 m
segments the same deformation is transmitted
to only 9 joints. Obviously, this decrease of the
number of the joints, unavoidably increases the
opening of each joint. Observe that the total
de-compression (of all joints) remains
practically the same : xtotal 1.5 m.
n insight into the aforementioned dynamic
readjustment of the tunnel segments can be
developed with the help of Fig 18, which
illustrates
the
history
of
longitudinal
deformations of six characteristic (Type B) joints
along the 70 m segment tunnel, subjected to
fault rupturing at Position 2. Initially, when only
the hydrostatic compression has been applied
(Step 1), all joints are compressed by exactly
the same amount, reaching x 28 cm.

decompression and re-compression. The total


dynamic plus static compression does not, in
any case, exceed the acceptable limits of the
Type B gasket : x (max) (29 cm). More
importantly, the gasket remains always
compressed, maintaining its water tightness :
(x)max (10 cm), in all joints. A redistribution of
joint deformation is readily observed :
overstressed joints tend to de-compress;
understressed to re-compress, so that the
safety of the immersed tunnel improves after
strong shaking.
This is a fundamental conclusion, observed
in all of the parametric cases analysed.
Moreover, it is completely explainable
theoretically. Its practical significance and its
generality should not be underestimated : a
detrimental static de-compression of a few of
the joints (arising from a rupturing fault, or from
differential settlement, or from any other case)
will be relaxed from the tunnel oscillations
during a strong seismic shaking.
In the transverse direction, the deformation
of the joints is restrained by the shear-key
"allowance" of only 5 mm. Deformations greater
than this value imply elastic deformation of the
shear key concrete. The terminal joints
experience the greatest relative displacement
y, with the maximum value being practically
equal to the shear-key allowance ( 5 mm).
Near the center, the transversal displacement
never exceeds 2 mm. These displacements are
mainly due to rotation between segments. Near
the two ends of the tunnel, where the rotation is
restrained by the stiff connection to the bored
approach tunnels, the shear keys experience
closure, without significant gapping taking
place. The transversal deformation is hardly
affected by the gasket type; it depends mainly
on the shear-key allowance.
DYNAMIC RE-ADJUSTMENT OF TUNNEL
SEGMENTS : A HEALING PROCESS

As already cited, during the asynchronous


dynamic oscillation of the tunnel, a
redistribution of total deformation among joints
is observed. This section further discusses this
interesting phenomenon.
As depicted in Fig 17, the application of the
faultinduced dislocation leads to a decompression of the immersion joints. When the
tunnel is at Position 2 relative to the
propagating fault rupture (center of the tunnel

22

40

40

(a)
Initial Hydrostatic Compression

Initial Hydrostatic Compression

30

x (m)

x (m)

30

(b)

20

10

20

10
Hydrostatic

Hydrostatic

Position 2

0
0

200

Fault
Position
1

Fault

0
400

600

800

1000

200

Horizontal distance d (m)


Hydrostatic

40

400

600

800

(c)

Fault

40

Hydrostatic

(d)

Hydrostatic

Fault

Fault

Initial Hydrostatic Compression

30

x (m)

x (m)

30

Initial Hydrostatic Compression

20

1000

Horizontal distance d (m)

10

20

10

Position 2

0
0

200

Position 2

0
400

600

800

1000

Horizontal distance d (m)

200

400

600

800

1000

Horizontal distance d (m)

Figure 17: Fault-induced decompression of the immersion joints : (a) 70 m segments with Type B gasket, at
Position 2 (max tensile displacement), (b) 70 m segments with Type B gasket, at Position 1
(max bending displacement), (c) 70 m segments with Type A gasket, at Position 2, and (d) 100
m segments with Type B gasket, at Position 2. 1

time (Step 1), joint 7 is significantly


decompressed to 12 cm : point a7 in the graph.
Now look into the hyperelastic load-deformation
curves of the two joints (Fig 5b). Joint 1, after
the application of the faultoriGinating
displacement remains sufficiently compressed
to lie on the stiffer hyperelastic regime, where
the stiffness is Khyper 50 MPa. By contrast,
joint 7, having decompressed significantly, lies
on the initial soft (elastic) branch, of Kel 3
MPa. As a result, during the dynamic shaking,
joint 1 will be an order of magnitude stiffer than
joint 7 ! Since the joints are connected in series,
the stiffer ones simply force the softer to
compress. In other words, over-compressed
joints (stiffer) force the under-compressed joints
(softer) to re-compress. Of course, such a recompression leads in turn to hyperelastic
stiffening, making the latter stiffer. Finally, an
equilibrium is reached, with all joint
compressions
being
substantially

Application of the faultinduced dislocation


(Step 2) leads to a differential de-compression
of the joints, as already explained in Fig 17. At
Step 3 the dynamic oscillation tends to
redistribute the joint compressions and relocate
the tunnel segments. Observe that joint 1,
which had experienced the least decompression, now experiences the greatest decompression during dynamic oscillation. On the
contrary, joint 7 at the center of the tunnel, with
the largest faultinduced de-compression, is recompressed the most during shaking.
This surprisingly favorable performance is
partly attributable to the hyper-elastic behaviour
of the immersion joints, combined with the
ability of the tunnel segments to slide over the
seabed. Let us now look into the performance
of joints 1 and 7 in more detail. As depicted in
Fig 18, the faultinduced stressing only slightly
decompresses the rubber gasket of joint 1 (from
28 to 26 cm) : point a1 in the graph. At the same

23

safely resist :
a normal earthquake fault rupture with a
dislocation (offset) of 2 meters in the
basement rock, 800 m underneath the
tunnel.
a subsequent strong ground shaking
arising from a different significant
seismic fault not crossing the site, but
producing peak acceleration of at least
0.50 g.
The cumulative effect of the two events,
even in the worst sequence : fault
rupture followed by shaking.
[2] The initial hydrostatic compressive force is
independent of segment length, or of total
number of the joints. Increasing the total
number of the joints (or decreasing the
segment length) leads to increased total
initial
hydrostatic
compressive
deformation of the tunnel. To ensure

homogenized. At the end of shaking, the


stiffer joint 1 exhibits a residual de-compression
at x 20 cm (point b1 in the graph), while the
softer joint 7 is re-compressed to x 13 cm
(point b7). It must be pointed out that such a
behaviour would hardly be possible if it were
not for the tunnel sliding : If the segments were
fixed to the seabed, then the stiffness of the
joints would play no, or very minor, role.
SUMMARY CONCLUSIONS

Several conclusions of practical significance


can be drawn from the presented study
(including additional parametric results not
shown here for the sake of brevity) :
[1] A properly designed immersed tunnel
(suitable thick elastic gaskets, small length
of segments, shear keys with sufficient
allowance, un-stressed tendons) can

30

Step 0 :

Step 1 :

Step 2 :

Hydrostatic
Compression

Fault-induced
Displacement

Dynamic
Oscillation
1
a1

11

25
2
b1

20

3
11

15

b7

9
a7

10
Load (MN/m)

1
Load (MN/m)

x
(cm)

Khyper

4
2
0

b1

Kel
0

10

20

a1
30

Khyper

4
2
0

40

Kel
0

x (cm)

0
-6

-4

-2

a7 b7
10

20

30

40

x (cm)

10

t (sec)

Figure 18: Longitudinal joint deformation for the 70 m segment tunnel, equipped with the thick Type B
rubber gasket, and 5 mm shear-key allowance, at Position 2 (maximum tensile deformation)
relative to the fault rupture. While the fault-induced displacement in Step 1 opens the joints, the
asynchronous dynamic shaking tends to relocate the segments and allow for a redistribution of
the gasket deformations : observe the homogenization of the residual compression of the joints.

24

[3]

[4]

[5]

water-tightness, the initial compression of


the Gina gaskets must be significant.
When subjected to faultinduced tensile
displacements, their ability to remain
water-tight is highly dependent on the
extent of the initial compression to avoid
developing net tension. On the other
hand, the additional dynamic compression
must not exceed the capacity of the joints.
The dynamicallyinduced longitudinal
deformation of the immersion joints
depends on the segment length and the
thickness of the Gina gasket. Increasing
this length unavoidably increases both the
faultinduced
and
the
dynamic
deformation of the immersion joints. Since
the tunnel segments are significantly
stiffer than the Gina gaskets, they tend to
behave as rigid blocks and most of the
imposed deformation is absorbed in the
joints. Obviously, decreasing the number
of the joints, increases their deformation.
Increasing the thickness of the Gina
gasket leads to greater initial hydrostatic
compression : from 17 cm for the Type A
gasket, to 28.5 cm for the Type B. Since
the tectonic deformation is mainly tensile,
this increase of the initial compressive
deformation leads to higher margins of
safety. With seismic shaking, the dynamic
(additional) re-compression and decompression of the gaskets is not so
sensitive to the gasket thickness.
During (asynchronous) dynamic shaking,
tunnel segments tend to re-adjust
themselves by sliding longitudinally over
the seabed. This is accompanied by a
redistribution of the longitudinal joint
deformation leading to a more uniform
profile of compression along the axis. This
surprising
healing
behaviour
is
attributable (to a substantial degree) to the
hyperelasticity of the rubber gaskets,
combined with the capability of the
segments to slide on the seabed. Thus,
while the fault displacements lead to a
detrimental de-compression different from
joint to joint, upon subsequent seismic
shaking
the
most
seriously
decompressed joints become more flexible
compared to joints with minor decompression. Since the joints are
connected in series, the stiffer ones will

force the softer to re-compressa


beneficial effect of great significance.
ACKNOWLEDGEMENTS

The authors would like to acknowledge the


financing of this research project by the Greek
Railway Organization (OSE). The help over the
years by Dr. Takashi Tazoh of Shimizu
Corporation who exposed us in the Japanese
technology was invaluable. We would also like
to thank Messieurs Hans van Italie, Hendrik
Postma, Gerard H. van Raalte, P. van der Burg,
and Royal Boskalis S.A. for kindly offering their
critical comments and suggestions on different
construction-related issues.
REFERENCES AND BIBLIOGRAPHY
Anastasopoulos I., & Gazetas G. (2007),
Foundation-Structure Systems over a Rupturing
Normal Fault : I. Observations after the Kocaeli
1999 Earthquake. Bulletin of Earthquake
Engineering, Vol. 5, No. 3.
Anastasopoulos I., & Gazetas G. (2007b),
Behaviour of StructureFoundation Systems
over a Rupturing Normal Fault : II. Analyses,
Experiments, and the Kocaeli Case Histories,
Bulletin of Earthquake Engineering, Vol. 5, No. 3.
Anastasopoulos, I., Gerolymos, N., Drosos, V.,
Kourkoulis, R, Georgarakos,P., and Gazetas, G.
(2007), Behaviour of Deep Immersed Tunnel
in
Strong
Seismic
Shaking, Journal of
Geotechnical
and
Geoenvironmental
Engineering, ASCE, Vol. 133.
Anastasopoulos I., Gazetas G., Bransby M.F.,
Davies M.C.R., and El Nahas A. (2007), Fault
Rupture Propagation through Sand : Finite
Element Analysis and Validation through
Centrifuge
Experiments,
Journal
of
Geotechnical
and
Geoenvironmental
Engineering, ASCE, Vol. 133.
Bickel, J.O., Tanner, D.N. (1982), Sunken tube
tunnels, in : Bickel, J.O., Keusel, T.R. Eds.,
Tunnel Engineering Handbook, Chapter 13, Van
Nostrand Reinhold, New York, pp. 354-394.
Bray, J.D., Seed, R.B., Cluff, L.S., and Seed, H.B.
(1994), Earthquake Fault Rupture Propagation
through
Soil,
Journal
of
Geotechnical
Engineering, ASCE, Vol. 120, No.3, pp. 543-561.
Bray, J.D., Seed, R.B., and Seed, H.B. (1994),
Analysis
of
Earthquake
Fault
Rupture
Propagation through Cohesive Soil, Journal of

25

Geotechnical Engineering, ASCE, Vol. 120, No.3,


pp. 562-580.

Poulos, H.G., & Davis, E.H. (1974), Elastic Solutions


for Soil and Rock Mechanics, J. Wiley & Sons.

Davis, R.O. & Selvadurai, A.P.S. (1996), Elasticity


and Geomechanics, Cambridge University Press.

Power, M.S., Rosidi, D., Kaneshiro, J. (1996),


Screening, evaluation, and retrofit design of
tunnels, Report, National Center for Earthquake
Engineering Research, Buffalo, New York.

Davies, M.C. & Bransby F. (2004), Centrifuge


experiments for fault rupture propagation,
University of Dundee Research Report.

Schnabel, P.B., Lysmer, J., Seed, B.H. (1972),


SHAKE : a computer program for earthquake
response analysis of horizontally layered sites,
Report no. EERC 72 /12, University of California,
Berkeley, CA, USA.

EC8 (2002) Eurocode 8 : Design of structures for


earthquake resistance. Part 2 : Bridges,
European Committee for Standardisation (CEN).
Gazetas, G. (1983), Analysis of machine foundation
vibrations: state of the art, Soil Dynamics and
Earthquake Engineering, Vol. 2, No. 1, pp. 243.

St. John, C.M. and Zahrah, T.F. (1987), Aseismic


design of underground structures, Tunnelling and
Underground Space Technology, Vol. 2, No 2,
pp. 165 -197.

Gazetas, G. (1991) Foundation Vibrations,


Foundation Engineering Handbook, 2nd edition,
ed. H.Y.Fang, Van Nostrand Reinhold.

Tselentis, A. (2004), Estimation of Seismic Hazard


and Seismic Deformation of the Rion-Antirrion
Straits, Technical Report, Earth Research Ltd,
July 2004.

Gerolymos, N., and Gazetas, G. (2005),


Constitutive Model for 1-D Cyclic Soil Behaviour
Applied to Seismic Analysis of Layered Deposits,
Soils and Foundations, Vol. 45, No. 3, June.

Veletsos, A.S., and Tang, Y. (1990), Deterministic


assessment of effects of ground motion
incoherence, Journal of Engineering Mechanics,
ASCE, Vol. 116, No. 5, pp. 1109-1124.

Ingerslev, C., and Kiyomiya, O. (1997), Earthquake


Analysis, Immersed and Floating Tunnels
Working Group Report, ITA, Vol. 2, No. 2, pp.
76-80.

Vrettos, Ch. (2005), Design issues for immersed


tunnel foundations, Proc. 1st Greece-Japan
Workshop on Seismic Design, Observation, and
Retrofit of Foundations, G. Gazetas, Y. Goto, T.
Tazoh (editors), pp. 257-266.

Ishibashi, I., and Zhang, X. (1993) Unified Dynamic


Shear Moduli and Damping Ratios of Sand and
Clay, Soils and Foundations, Vol. 33, No. 1, pp.
182-191.
Johansson, J., Konagai, K. (2004), Fault Induced
Permanent Ground Deformations Simulations
and Experimental Verification, Proc. 13th World
Conference on Earthquake Engineering.
JSCE (1975), Specifications for Earthquake
Resistant Design of Submerged Tunnels,
Japanese Society of Civil Engineers, Tokyo.
JSCE (1988), Earthquake Resistant Design for Civil
Engineering Structures in Japan, Japanese
Society of Civil Engineers, Tokyo.
Kiyomiya, O. (1995) Earthquake-resistant Design
Features of Immersed Tunnels in Japan,
Tunneling and Underground Space Technology,
Elsevier, Vol. 10, No. 4, pp. 463-475.
Kuesel, T.R. (1969). Earthquake Design Criteria for
Subways. Journal Structural Division, ASCE
ST6, pp. 1213-1231.
Nakamura, S., Yoshida, N., Iwatate, T. (1996),
Damage to Daikai Subway Station During the
1995 Hyogoken-Nambu Earthquake and its
Investigation, Japan Society of Civil Engineers,
Report on Hyogoken-Nambu Earthquake, pp.
287-295.

26

Seismic Triggering, Evolution, Deposition, and Retaining


of Rapid Landslides
Nikos Gerolymos, George Gazetas, and Ioannis Vardoulakis
National Technical University, Athens, Greece

Abstract
A mathematical model is developed for the dynamic analysis of earthquaketriggered rapid
landslide, considering two mechanically coupled substructures: (a) the accelerating
deformable body of the slide, and (b) the rapidly deforming shear band at the base of the
slide. The main body of the slide is considered as an one-phase mixture of Newtonian
incompressible fluids and Coulomb solids sliding on a plane of variable inclination. The
evolution of the landslide is modeled via an extended SavageHutter model coupled with the
MohrCoulomb sliding law for the frictional deformation of the material within the shear band.
The capability of the model is tested through analysis of the HigashiTakezawa landslide,
triggered by the 2004 Niigata-Ken Chuetsu earthquake. The mechanism of material softening
inside the shear band responsible for the 100mrunoff of the landslide, is described by a set
of equations for grain crushinginduced porewater pressures (Gerolymos and Gazetas,
2007). Then, the developed model for landslide kinematics is appropriately modified to
include the dynamic interaction between a rapidly moving landslide and a retaining piled
wall. Characteristic examples highlight the influence of both the bed inclination and the
reaction thrust exerted by the sliding mass on the wall, on the de-acceleration and finally
termination of the landslide. A limiting equilibrium approach to the problem is also presented,
resulting in design diagrams of the maximum impact velocity the wall experiences as a
function of the local landslide thickness and structural strength of the wall.

INTRODUCTION

by the inertial loading it imposes, or by causing


a loss of strength in the slope materials which
may result to catastrophic landslides, as
illustrated in Fig 1.
Analysis of rapid longrun out landslides is a
difficult challenge in hazard studies because
they endanger areas situated far from the
landslide source. Prediction of run-out distance
is a key requirement for delineation of the
hazard zone. Moreover, prediction of dynamic
parameters such as flow velocity and depth are
necessary for the design of protective
measures.
There is a variety of landslide countermeasures used to control the movement of
debris slide masses without attempting to
stabilize them (Baldwin et al., 1987; Hungr et al.
1984). Among them catch fences and trap
ditches are widely used to trap rockslide debris
while reinforced concrete retaining structures
are appropriate for debris flows and landslides.

Landslides are a serious geologic hazard


causing severe damage to structural facilities
and numerous of deaths and injuries each year.
Some landslides move slowly and induce a
limited damage, whereas others move rapidly
transporting downstream sometimes huge
volumes of sediments, covering large run-out
distances and destroying everything in their
passage. Gravity is the force that drives the
landslide movement. Factors that allow the
force of gravity to overcome the frictional
resistance of the soil and to trigger the
landslide, include: saturation by water (e.g. after
periods of heavy rainfall or rapid snow melt),
steepening of slopes by erosion or construction,
frictional softening due to earthquake induced
pore water pressures. Among the possible
causes of landsliding initiation, earthquake
shaking is of particular interest. A seismic
shaking can cause a slope to become unstable

27

Fig 1: Photos of characteristic earthquakeinduced catastrophic landslides. (left) landslide triggered by the
2001 El Salvador Mw = 7.1 earthquake. The landslide mass buried hundred of residences and accounted for
over half of the nearly 700 earthquake victims (the photograph was produced by Corbis Coorporation). (Top
right) landslide triggered by the 2004 Niigata-Ken Chuetsu MJMA = 6.8 earthquake (by Koichi Kamoshida/Getty
Images). (Bottom right) Massive landslides on the outskirts of Muzaffarabad, triggered by the 2005 Pakistan
Mw = 7.6 earthquake (by Image courtesy DLR, Kathryn Cramer, Google Earth).

a versatile and economical approach and has


been proved to be effective for the backanalysis of landslides with large run-out
distance. Nevertheless, it has the following
deficiencies: It provides limited information on
the evolution of the landslide, and it is not
capable of reproducing the mechanisms of
deposition of the slope material.
On the other hand, the deformability of the
sliding mass is fully considered by an elastoplastic solid model. A variety of constitutive
laws of varying degree of accuracy are
available to reproduce the nonlinear stress
strain behaviour of the slope material. Complex
phenomena such as material softening and loss
of strength due to porewater pressure
development can be readily modeled within the
framework of elasto-plasticity. Theory of elastoplasticity has been extensively used in
predicting the failure mechanism of a potential
landslide. Sophisticated methods based on
gradient and non-local constitutive models are

Evidently, numerical modeling of landslide


retaining wall interaction systems could serve
as a tool for evaluating landslide hazard.
The models for the analysis of earthquake
triggered landslide could be classified into three
broad categories: (a) Newmark sliding block
model and its extensions, (b) Elasto-plastic
solid models, and (c) Rheologicaldepth integrated
models. Each has advantages and limitations.

The former has been widely used in


geotechnical engineering for deformation
analysis of earthquakeinduced landslides. The
sliding block method assumes the potential
mass to be rigid, even though conditions for
actual slopes certainly vary from this
assumption. The accuracy of a sliding block
analysis depends strongly on the yield
acceleration. That is the threshold acceleration
at initiation of sliding. Obviously, the model
predicts zero permanent slope displacement if
earthquake induced accelerations never exceed
the threshold value. The sliding block model is

28

model (Laigle and Coussot 1997), and the


quadratic shear stress model (O Brien and
Julien 1985). What these models have in
common is that the total dynamic frictional
resistance is the resultant of a hysteretic
(Coulomb) and of a viscous (velocity
dependent) component.
On the other hand, in a variable density
mixture, the limitations mentioned above vanish
allowing for a wider choice of rheological
models. The coarser sediments settle even
though the remaining mixture continues to flow
downstream. In a two-phase mixture model,
termination of the landslide could be achieved
even for minimal frictional resistance and nonzero bed inclination.
A considerable amount of studies has been
dedicated the last decade to the development
of rheological depthintegrated models of
varying degrees of accuracy. Sophisticated
numerical techniques have been utilized to
solve the governing equations even for arbitrary
topographies. As to the authors knowledge
none of those studies address the interesting
issue of earthquaketriggered landslides, while
most of them suggest simple hysteretic frictional
laws (e.g. constant Coulomb friction) insufficient
to reproduce complicated material softening
behaviour which usually accompanies the
deformation of a shear band.
In this paper the governing equations of
landslide motion, originally proposed by Savage
and Hutter (1989) and extended by Iverson
(1997) and Gray (1999), are reformulated to
account for inertial force due to seismic loading.
The landslide mass is considered as an onephase mixture of incompressible fluids and
solids sliding on a plane of variable inclination.
The governing equations are coupled with a
BoucWen type constitutive model (Gerolymos
and Gazetas 2005, Gerolymos et al. 2007) for
the hysteretic stressdisplacement behaviour of
the shear band in cyclic loading, in conjunction
with the MohrCoulomb failure law for the
frictional resistance, and the quadratic Chezy
law for the viscous (turbulent) shear resistance.
The proposed model is applied to the
analysis of HigashiTakezawa landslide,
triggered by the 2004 Niigata-Ken Chuetsu
earthquake. The landslide mass covered a
distance of 100 m, filled a valley and stopped a
river flow forming a large natural reservoir. The
surprisingly large and rapid runoff of the soil
mass motivated several researchers (Kokusho

capable of simulating the formation and


development of shear zones with great
accuracy and thus improving the modeling of
failure mechanisms. However, it has not yet
been examined as to whether those models are
able to describe postfailure behaviour.
Moreover, the field equations of elasto-plastic
analysis are generally formulated by coupled
equations of soil and pore water based on the
solid mechanics within the framework of
infinitesimal strain. However, the runoff distance
of a landslide usually ranges from a few meters
to a few hundred meters. And there are more
limitations: (i) Due to strain localization
processes at the base of the potential failure
mass, the landslide behaves as a nearly rigid
body after its initiation. And (ii) Erosion and
deposition processes during the landslide can
hardly be reproduced by elasto-plastic models.
Rheological models overcome most of the
limitations mentioned above. These models
assume that the sliding mass behave as a liquid
mixture of interacting fluids and solids. Solid
and fluid constituents obey mass and
momentum balances, which are summed and
depthintegrated to yield equations that
describe shallow flow of the mixture.
Rheological depthintegrated models were
originally developed to simulate debris flow, but
they have been also successively used in
modeling
post-liquefaction
behaviour
of
granular soil. The mixture can be considered
either as an one-phase fluid of constant density,
or as a two-phase mixture of variable density
composed by granular material immersed in an
interstitial fluid.
A constant density fluid model cannot,
however, simulate the erosion/deposition
process in which the coarser sediments settle in
the upper part of the alluvial fan or near
obstacles in the river bed. Moreover, it is well
known that erosiondeposition mechanism
significantly contributes to the de-acceleration
and termination of the landslide irrespectively of
the topographic inclination and the frictional
resistance on the sliding base. Evidently, the
predictive power of one-phase fluid models is
strongly influenced by the choice of an
appropriate frictional constitutive model in which
a minimum value for the soil shear strength is
assumed. Such models are: the Bingham
model (Fraccarollo 1995, Jan 1997, Jin and
Fread 1997, Chen and Lee 2002, Hadush et al.
2000, Uzuoka et al. 1998), the HerchelBulkley

29

the wall on the sliding mass, on the deacceleration and termination of the landslide.
Finally, a limiting equilibrium approach to the
problem is also provided, resulting in diagrams
of the maximum impact velocity the wall
experiences as a function of the local landslide
thickness and structural strength of the wall,
which can be readily utilized in the design of the
retaining structure.

and Ishizawa, 2005; Tsukamoto and Ishihara,


2005, Sassa et al., 2005) to study the Higashi
Takezawa
landslide,
providing
different
interpretations of the sliding process. The
questions to be answered arose on: (a) the
exact position of the sliding surface, and (b) the
mechanism of material softening behind the
accelerating landslide movement. Laboratory
tests on soil samples taken from the site of the
slip surface indicated undrained friction angles
larger than the slip inclination (Sassa et al.,
2005). Moreover, the sliding material consisting
of silt to dense silty sand was not susceptible to
liquefaction (Kokusho and Ishizawa, 2005).
However, experimental evidences revealed the
mechanical instability of the shear band
material due to grain crushing. For this, a
constitutive model (Gerolymos and Gazetas
2007) is utilized that mathematically interprets
the concept of high porewater pressure
generation by grain crushing along the sliding
surface. The constitutive model for shear band
behaviour coupled with the depthintegrated
model for landslide kinematics, reproduces
satisfactorily the field observations.
The developed depthintegrated model is
extended and further utilized to model the
interaction of a retaining piledwall with a
rapidly deforming landslide. The piledwall is
represented as a flexible beam, while the wall
soil interface is replaced with continuously
distributed nonlinear dashpots. Computation of
the impact forces exerted by the flowing soil
mass on the wall is based on analytical
formulae proposed by Morison (1950) and
Armanini and Scotton (1993). The bending
moment of the wall section is a nonlinear
function of its curvature obeying a criterion of
the extended BoucWen type (Gerolymos and
Gazetas, 2005).
The derived system of non-linear differential
equations is solved numerically applying an
explicit central difference numerical technique
in conjunction with a shockcupturing scheme
to avoid spurious oscillations originating from
abrupt variations of the model properties. The
capability of the developed model to analyse
the evolution of a landslide on a bed with
varying inclination and its interaction with a
retaining structure constructed in front of the of
the landslide, is investigated through a
parametric study. Valuable conclusions are
drawn regarding the role of both the bed
inclination and the reaction thrust exerted by

THE MODEL : EQUATIONS AND


PARAMETERS
Problem Definition

The problem studied is that of a finite


moving soil mass assembled by a number of
columns in contact with each other (Fig 2). The
columns are free to deform but retain fixed
volumes (constant density ) of solidfluid
mixtures during their movement down a slope.
The evolution of the mixture is considered to be
onedimensional with no aggradation or
degradation processes and with uniformly
distributed (depthintegrated) velocity along
each column. At the base of the sliding mass
we assume a shear band of infinite length and
of thickness db (Fig 2), subjected to an
acceleration time history uniformly imposed
along the entire length of the sliding base. The
material inside the shear band is assumed to be
free of porewater pressures. The shear zone
deforms with a velocity (t, x) equal to that of
the sliding mass.
Applying the mass and momentum
conservation laws and using Eulerian
description of motion, a system of two partial
differential equations are obtained:

h
h

+
+h
=0
t
x
x

(1)

and


(h x )
g

+ Td Tr T f = h
+
+
x
t
x
t

(2)
h is the depth in the z direction normal to the
bed (Fig 2), is the depthaveraged velocity in
the x direction parallel to the base of the
landslide, g is the seismic acceleration

30

(x)

h(x,t)
(x,t)

bedrock

dx
A(t)hdx

x +

gcos
Tr + Tf

A(t)

x
dx
x

Td

Fig 2: 1 dimensional depthintegrated model for the analysis of earthquakeinduced landslide evolution.
Stress equilibrium is referenced to a local coordinate system that is fitted to the underlying topography.

imposed at the base of the landslide parallel to


the dip direction of the sliding surface. Note that
spatial variability of the seismic motion has
been neglected, and the same acceleration
time history is imposed instantaneously along
the entire length of the sliding surface. Td is the
gravitational driving force acting on the
landslide mass

Td = g h sin

in which m is the cyclic shear resistance


mobilized along the shear band. A detailed
description of this term will be provided below.
Tf is the turbulent resisting force at the base of
the slide, represented by the quadratic Chezy
constitutive law

T f = T 2 sgn( )

(3)

in which is the turbulent (or hydraulic) friction


coefficient relating the viscous shear stress with
landslide velocity. Exploiting experimental
results from 1g models, Ancey et al. showed
that at small flow velocities (quasi-uniform flow
regime) is a linear function of the landslide
depth according to

in which g is the gravitational acceleration and


is the local bed slope. Tr is the resisting force
due to hysteretic (Coulomb) friction at the bed
influenced by bed curvature (Gray et al. 1999,
Iverson and Delinger 2001)

2
m
Tr = 1

g
x
cos

(5)

(4)

T = h

31

1 1

40 D g

(6)

in which Dg is the mean grain diameter. For


larger velocities, however, corresponding to
Froude numbers Fr > 1 (splashing flow regime),
assumes a constant depth-independent
value. Blagovechshenskiy et al. (2002), backanalysing measured field avalanche motion and
assuming that the hydraulic friction coefficient
is depth-independent, showed that is a
decreasing function of velocity with a range of
values from 0.01 to 0.10 for velocities between
3 m/s and 35 m/s. In general varies
significantly from avalanche to avalanche
(Hutter and Greve, 1993) and its contribution to
debris flow evolution can hardly be
distinguished from that of Coulomb friction.
Zwinger (2000) studied the Madlein avalanche
of 1984 and conjectured that reasonable
agreement with the observed deposition could
only be obtained when the sliding law was
Coulomb-type at low velocities and viscoustype at large velocities. Moreover, all the
aforementioned conjectures concern debris
flows with aspect ratios between 102 to 103
rather than landslides which are usually
characterized by aspect ratios of the order of 10
and thus exhibiting a completely different
rheological behaviour.
In Eqn(2), x is the average along the
depth of the sliding mass longitudinal normal
stress due to elongation or compression of the
soil mass in the x direction. The longitudinal
normal stress is assumed to be a combination
of a lithostatic (depthdependent) term and a
strain rate dependent term

x =

1
s g h cos d h
x
2

in which s , and d are the lithostatic and


viscous coefficients, respectively. is the
aspect ratio of the sliding mass, designated as

typical thichness [H c ]
typical length tangential to the bed [Lc ]

(8)
Pudasaini and Hutter (2003) suggest that
this ratio is of the order of 10-3 to 10-1 for typical
granular avalanches, with smaller values
corresponding to debris flow and larger ones to
landslides. Eqn(7) can be rewritten in the
following form

x =

1
K dyn g h cos
2

(9)

in which Kdyn is the dynamic lateral earth


pressure coefficient determined as

K dyn = s

2 d
g cos x

(10)

Kdyn varies in the following range of values

K a K dyn K p

(11)

Ka and Kp are the active and passive lateral


earth pressure coefficients, respectively,
derived from Coulomb failure theory and
determines the magnitude of the horizontal
normal stress at yielding, depending on whether
a soil column is expanding or contracting. For a
cohessionless (granular) soil, the following
expressions for K have been proposed by
Iverson and Delinger (2001):

(7)

1 + 1 cos 2 (1 + tan 2 )
1 ,
Kp = 2
cos 2

K =

2
2
K = 2 1 1 cos (1 + tan ) 1 ,
a
cos 2

in which is the internal friction angle of the soil


mixture, and the friction angle at the base of
the sliding mass. Similar relationships for the

0
x
(12)

>0
x

lateral earth pressure coefficient, based


however on different assumptions, have been
also proposed by Pudasaini and Hutter (2003).

32

1
d
=
1
du
uy

Eqn (12) is not valid for cohesive soils.


According to Rankines theory for lateral earth
pressure, negative values of K are also possible
when cohesion is taken into account.
The first term in the righthand side of Eqn
(7) governs the rheological behaviour of the
sliding mass. A typical range of values for the
lithostatic coefficient s is between 0.01 and 1,
with larger values corresponding to deformation
of fluidlike mixture. For the special case s = 0,
the sliding mass behaves as a rigid body and
Eqn (6) vanishes to the well known Newmark
sliding block model.
The second term on the righthand side of
Eqn (7) plays the role of a damping factor for
the surface waves emanating from the frontal
edge of the landslide and propagating along its
crest. These waves could be generated due to
either (i) elongationcompression of the soil
mass in the direction of flow, or (ii) abrupt
variations of bed inclination, or (iii) impact of the
avalanche front (shock wave) on a retaining
structure. A typical range of values for the
damping coefficient d is between 1 and 10,
with smaller values representative of fluidlike
mixture (less damping), and larger values
corresponding to solidlike mixture.

[b + (1 b ) sgn( )] }

(14)

in which uy is a parameter signaling yielding in


the soil (a rigid plastic behaviour is
approximated by assuming a very small value
of uy , say less than < 10-3 m). uy is defined as
the ratio of the ultimate shear strength y to the
shear modulus G of the soil, multiplied by the
shear band thickness db. However, as the shear
band thickness is considered to be zero in our
problem, uy can be calculated alternatively from
a direct or from a ring shear test. n and b are
dimensionless quantities that control the shape
of the hysteresis loop. Eqn (14) is obviously of a
hysteretic rather than a viscous type. Hence, its
solution is not frequency dependent.
Parameter n governs the sharpness of the
transition from the linear to the nonlinear range,
during initial virgin loading. It ranges from 0 to
, with elasticperfectly-plastic behaviour
practically achieved when n takes values
greater than 10. Values of n between 0.6 and 1
have been found to better fit experimental
results (Gerolymos and Gazetas, 2005).
Parameter
b
controls
the
shape
of
unloadingreloading curve. Its range of values
is between 0 and 1. When b = 0.5 the stiffness
upon loading reversal equals the initial tangent
stiffness, and the Masing criterion for
loadingunloadingreloading arises. For more
details on Eqn (14) and calibration of the
associated parameters, the reader is referred to
the recent publication of Gerolymos and
Gazetas (2005).
The shear strength y is given by Coulombs
friction law in conjunction with the hysteretic
BoucWen model (Gerolymos and Gazetas,
2005;
Gerolymos
and
Gazetas
2007;
Gerolymos et al., 2007):

Equations for Frictional Behaviour of the Shear


Band

A versatile onedimensional macroscopic


model is utilized to describe the shear stress
strain relationship inside the shear band. The
model is capable of reproducing an almost
endless variety of stressstrain forms,
monotonic as well as cyclic. Based on the
original proposal by Bouc (1971) and Wen
(1976), the model was extended by Gerolymos
& Gazetas (2005) and applied to cyclic
response of soils. A simple version of the model
is briefly outlined here.
The mobilized shear stress inside the shear
band is expressed as:

m = y

y = ( n p )

(15)

in which p is the excess pore water pressure,


is the mobilized friction coefficient, expressed in
terms of the friction angle at the base of the
sliding soil mass

(13)

in which y is the ultimate shear strength of soil,


and is a dimensionless hysteretic parameter
controlling the nonlinear response; it is
expressed with the following differential
equation:

= tan
and

33

(16)

n = g h cos

silt encountered at the head scarp of the


Terrano landslide was well weathered and soft.

(17)

is the total stress normal to the shear band.


THE HIGASHITAKEZAWA LANDSLIDE
Background

The main body of the landslide is indicated


in the plan of Fig 3 deduced from an air borne
laser scanning survey (Sassa et al., 2005)
carried out three days after the earthquake. A
crosssection of the landslide is also depicted
in Fig 3. The gentle slope inclination before the
head scarp reveals that the landslide was a
reactivation of a previous one. The landslide
involved a soil volume of about 1 200 000 m3
(Kokusho and Ishizawa, 2005). The maximum
dimensions in plan were about 300 m width and
250 m length (Kokusho and Ishizawa, 2005),
and the maximum thickness was about 40 m
(Sassa et al., 2005). The landslide mass moved
rapidly around 100 m, and hit the opposite bank
of Imokawa river (Sassa et al., 2005). A part of
the sliding mass spread across the road and hit
a school. From the head scarp of the landslide,
consisting of a rather impermeable stiff
siltstone, the inclination angle of the sliding
surface was estimated to be approximately 20o
(Sassa et al., 2005; Kokusho and Ishizawa,
2005).
A schematic geological section of the
landslide area is shown in Fig 3. The subsoil is
essentially constituted of a Neogene formation,
consisting of sandstone (the main body of the
landslide) underlain by siltstone. The terrace
along the river and below the toe of the
landslide consists of marine sand from the
Tertiary period (Sassa et al., 2005). The
groundwater flow over the siltstone layer, lead
Sassa et al. (2005) to assume the existence of
a thin silt layer between the sandstone and the
siltstone, due to weathering of the siltstone.
Although, this silt layer was not detected at the
head scarp, the assumption of Sassa et al.
(2005) was reinforced from field investigation of
the head scarp of the Terrano landslide located
in the vicinity of HigashiTakezawa and near
the Immokawa river. The Terrano landslide
(Sassa et al., 2005), was also triggered by
NiigataKen Chuetsu earthquake, and had the
same subsoil and groundwater conditions. The

Fig. 3. The Higashi Takezawa landslide: (a) plan


view, and (b) cross section (Sassa et al., 2005)

Water seepage observed on the head scarp


of the landslide three days after the earthquake
suggests that the water table was located well
above the sliding plane. No precipitation was
observed during a period of several days
preceding the earthquake, which occurred
during the rainy season.
The grain size distribution of the sand
involved in the sliding surface of the Higashi
Takezawa landslide is illustrated in Fig 4, along
with that of the Terrano silt which is considered
to be representative of the HigashiTakezawa
one. The strength properties of the soils under
consideration were obtained from consolidated
drained and undrained high speed ring shear
tests (Sassa et al., 2005). The undrained

34

friction angle of the sand was found to be 36.9o,


while the residual friction angle of the Terrano
silt was 23.9o. However, one peculiar aspect of
the HigashiTakezawa sand is its mechanical
instability due to grain crushing. The cyclic
loading test, resulting in an apparent friction
angle of 3.3o, indicated that the Higashi
Takezawa sand is susceptible to grain
crushinginduced liquefaction.

In the limit of undrained loading conditions,


which is a reasonable assumption when the
shear band is deformed at a large velocity
(rapid landslide), and assuming that advection
dominates upon diffusion, Eqn (18) reduces to:

dB p
p
p
+ =
n 0
t
x
dt

Fraction finer by weight

Eqn (19) instead of (18) will be used in all


subsequent analyses. The parameter controls
the ultimate value of the porewater pressure.
The larger the value of , the higher the
asymptotic value of the porewater pressure.

0.8
0.6
0.4

Equations for grain crushing

0.2
0
0.001

(19)

0.01

0.1

As already discussed, the breakage


potential Bp is a measure of the evolution of the
particle size distribution curve with loading, and
hence of the amount of grain crushing. We
assume that the evolution of Bp with time is
governed by the following equation (Gerolymos
and Gazetas 2007):

10

particle diameter D (mm)

Fig. 4. Grain size distribution of the Higashi


Takezawa sand (black line) and Terrano silt (gray
line), after Sassa et al. (2005)
Equations for grain crushinginduced pore
water pressure

dB p
dt

The mechanism of porewater pressure


generation due to particle breakage is assumed
to be governed by the following equation
(Gerolymos and Gazetas, 2007):

= (B pl B p )

(20)

in which is the coefficient of grain crushing.


Increasing values of correspond to increasing
rates of porewater pressure rise. Bpl is the final
(after loading) breakage potential as computed
at the current time of loading, given by:

B p
p
p

p
+
=
n 0
cv (B p )
t
x
z
z
t
(18)

B pl =

in which Bp is the current value of the breakage


potential; cv and are the coefficients of
consolidation and porepressurebreakage,
respectively. Note that cv is a function of Bp. In
fact, cv decreases with decreasing particle size
and thus with particle crushing evolution. It is
also important to notice the advection term in
the lefthand side of Eqn (18) which is usually
ignored in a conventional soil consolidation
analysis, as the developed flow velocities are in
general extremely small. However, this is not
valid for a rapidly evolved landslide in which
transport of the excess porewater pressure
along the shear band is a significant
mechanism of the sliding process, and
therefore can not be neglected.

B p0

1 + S nb

(21)

in which Bp0 is the initial (before loading) value


of Bp, defined as (Hardin, 1985). The definition
of Bp0 is schematically illustrated in Fig 5. The
breakage number, nb, is expressed according
to:
2

nb =

hc
+ 0.3
(1 + e0 ) n s

(22)

in which hc and ns is the crushing hardness and


shape number of the particle, respectively; and
e0 is the initial void ratio of the particles mixture.
For more details on the definitions of those

35

occurs only after the yield surface has been


reached, by contrast to conventional (mass)
liquefaction in which degradation of shear
resistance initiates below the yield surface,
when the phase transformation line has been
reached.

parameters the reader is referred to the work of


Hardin (1985). In Eqn (21), S is the stress
loading factor. In an undrained cyclic simple
shear test S is given by the following equation:
3

0.3

(23)

Shear resistance (MPa)

1 + e0 n m

S = 9
800 h 2 p a n

For a given shear stress time history, Eqns


(13), (14), (18), and (20) form a system of highly
nonlinear partial differential equations with four
unknowns: the excess porewater pressure p,
the breakage potential Bp, the hysteretic
parameter , and the displacement u.

0.2

0.1

(a)
0
0

0.1

Excess pore-water pressure ratio r u

25

oa
din
g
el
Be
for

50

ter
loa
din
g

Bp0

Silt

Af

Percent finer by weight (%)

Evolution of grain size


distribution with loading

75

0
0.01

0.2

0.3

0.4

0.5

Effective normal stress (MPa)

100

0.074 0.1

10
1
Particle diameter D (mm)

100

1
0.8
0.6
0.4

(b)

0.2
0
0

10

15

20

25

30

20

25

30

t (sec)

Fig. 5. Definition of the initial breakage potential Bp0,


after Hardin (1985)
Shear displacement (m)

Calibration of model parameters

Calibration of the parameters for shear band


behaviour is achieved through numerical
simulation of undrained cyclic ring shear tests
conducted by Sassa et al. (2005). The shape
number, crushing hardness, and initial void ratio
were assumed to be ns = 25, hc = 2.4, and e0 =
0.6, respectively, while the initial breakage
potential was calculated from Fig 4 to Bp0 =
0.34.
Detailed
information
on
the
aforementioned parameters is given by Hardin
(1985).
The experimental results are reproduced in
Figs 6 in the form of time history of the
developed shear displacement, and plot of the
shear resistance versus effective normal stress.
The breakage and porepressure breakage
coefficients correspond to the analysis are =
0.05 and = 25. Note that the proposed model
is capable of reproducing the brittle behaviour
of a soil undergoing grain crushinginduced
porepressure. That is, loss of shear resistance

0.8
0.6
0.4

(c)

0.2
0
0

10

15

t (sec)

Fig. 6. (a) Computed stress path (the loading is


plotted with gray line), and time histories of (b)
excess porewater pressure ratio, and (c) shear
displacement of the undrained cyclic ring shear test
of the HigashiTakezawa sand, computed with the
proposed model. The circles correspond to the
experimental data of Sassa et al. (2005)
ANALYSIS OF THE HIGASHITAKEZAWA
LANDSLIDE

With the developed model for seismic


triggering and evolution of graincrushing
induced landslides, we analyse the case of
HigashiTakezawa. The parameters are: s = 2

36

t / m3, K = 0.5, = 1, d = 5, T = 0.01h, n = 3, b


= 0.5, uy = 10-3 m, = 0.75, ns = 25, hc = 2.4, e0
= 0.6, Bp0 = 0.34, = 25, and = 0.05. The
seepage force is ignored, since the actual level
of the water table during the earthquake is not
known. The actual seismic excitation exerted on
the landslide cannot be known in detail, as it is
influenced by many parameters such as the
geology, topography, site conditions and
distance from the fault.

and the upper part of the siltstone is


assumed to have remained intact.
(b) The shear band formed within an assumed
thin silt layer atop the siltstone, but the sand
is not susceptible to grain crushing.
(c) The shear band formed within the sand
layer, the sand is susceptible to grain
crushing, and the upper part of the siltstone
is assumed to have remained intact.
The results of the analysis for cases (a) and
(b) are shown comparatively in Fig 7 in the form
of time histories of relative shear displacement.
The maximum computed displacement at the
end of shaking for case (a) is 0.65 m, which is
by far smaller than that of 3.4 m for case (b).
These values of displacement suggest that the
existence of a thin silt layer atop the siltstone is
more crucial for triggering the landslide.
However, none of them could explain the
observed rapid and large run-out distance of
the landslide. It is therefore reasonable to
assume that grain crushinginduced pore
pressures could be a major destabilizing factor
for the landslide.
The results of the analysis for case (c) are
presented in Fig 8 in terms of snapshots of the
landslide evolution, and distributions of velocity
(Fig 9), excess porewater pressure ratio ru
(Fig 10), and breakage potential Bp (Fig 11),
along the sliding surface. The following
observations are worthy of note regarding the
response of the sliding wedge:
At the early stages of the seismic motion,
excess pore waterpressure due to particle
crushing is generated at the head of the wedge
and propagates rapidly towards its toe. In the
following few seconds the excess porewater
pressure ratio rises up very quickly reaching
values larger than 0.9 along the entire length of
the sliding surface (t = 12.5 sec blue line). At
this time, sliding originates at the head of the
soil wedge, and landsliding begins. It is very
interesting that triggering occurs almost at the
end of seismic shaking, when the motion has
essentially subsided, and not during the strong
seismic shaking as one would expect. This
implies that grain crushinginduced pore
pressure is a cumulative process and thus
depends strongly on the history of loading.
After its initiation the landslide moves rapidly
towards the riverbed, developing velocities
between 5 m/s and 16 m/s. Velocities with
smaller values concentrate on the rear of the

15

Acceleration (m / s )

10
5
0
-5

(a)

-10
-15
0

10

15

20

t (sec)

Shear displacement (m)

(b)
0
0

10

15

20

t (sec)

Fig. 7. (a) Input acceleration time history (NIG019


EW 2004 - PGA = 1.3 g) at the base of the landslide,
and (b) computed time histories of relative shear
displacements, for sliding surface: (i) within the sand
layer(no grain crushing is considered, maxu = 0.65
m) (black line), and (ii) within an assumed thin silt
layer at the top of the siltstone ( maxu = 3.4 m) (gray
line)

Therefore, we apply as excitation the EW


component of the record from the nearest (to
the landslide) observation station NIG019 at
Ojiya (PGA = 1.3 g), around 10 km west of the
HigashiTakezawa landslide and WNW 7 km
from the epicenter of the main shock (Sassa et
al. 2005). Three scenarios are studied
regarding the potential location of the sliding
surface and the susceptibility of sand to grain
crushing:
(a) The shear band formed within the sand
layer (i.e., in the main body of the landslide),
the sand is not susceptible to grain crushing,

37

120
90
60
30

t=5s

0
120 0

50

100

150

200

250

300

350

400

100

150

200

250

300

350

400

100

150

200

250

300

350

400

100

150

200

250

300

350

400

100

150

200

250

300

350

400

100

150

200

250

300

350

400

90
60

t = 12.5 s

30
0
120

50

Elevation (m)

90
60

t = 19 s

30
0
120 0

50

90
60

t = 22.5 s

30
0
120 0

50

90
60

t = 26 s

30
0
120 0

50

90
60
30

t = 30 s

0
0

50

Horizontal distance (m)


Fig. 8 Snapshot of the computed landslide evolution, for the case of HigashiTakezawa. The school is
illustrated with the gray box

38

1.2

25
20
15
10
5
0

t=5s

0.9
0.6
0.3
0

100

150

200

250

300

350

400

50
25 0
t = 12.5 s
20
15
10
5
0

100

150

200

250

300

350

400

50

100

150

200

250

300

350

400

50
25 0
t = 19 s
20
15
10
5
0

100

150

200

250

300

350

400

50

100

150

200

250

300

350

400

50
25 0 t = 22.5
s
20
15
10
5
0

100

150

200

250

300

350

400

50

100

150

200

250

300

350

400

50
25 0
t = 26 s
20
15
10
5
0

100

150

200

250

300

350

400

50

100

150

200

250

300

350

400

50
25 0
t = 30 s
20
15
10
5
0

100

150

200

250

300

350

400

50

100

150

200

250

300

350

400

150

200

250

300

350

400

50
1.2 0 t = 12.5
s
0.9
0.6
0
0.9

t = 19 s

0.6
0.3
0
1.2 0
0.9

t = 22.5 s

0.6
0.3
0
1.2 0
0.9

t = 26 s

0.6
0.3
0
1.2 0
0.9

t = 30 s

0.6
0.3
0
0

Velocity (m / s)

Excess pore water pressure ratio ru

0.3
1.2 0

t=5s

50

100

Horizontal distance (m)

Horizontal distance (m)

Fig. 910 Snapshots of the computed excess porewater pressure ratio along the sliding surface (left), and
distribution of the landslide velocity, for the case of HigashiTakezawa. The school is illustrated with the gray
box

motion towards the riverbed ( 12.5 sec < t <


22.5 sec), separation of the frontal part from the
main body of the landslide (at t = 22.5 sec), and
deposition and deceleration ( t > 22.5 sec).

landslide, while those with larger values are


mostly at the front which essentially governs the
race of the entire landslide. At t = 22.5 sec the
sliding soil mass enters the riverbed while at
this time the frontal part of the landslide
detaches from the main body, spreads across
the river, hits the opposite bank with a velocity
of 25 m / s (at t = 26.5 sec), and finally reaches
the school at t = 30 sec. Following this frenetic
motion of the detached frontal part, the main
body of the landslide accumulates inside the
riverbed forming a natural reservoir which
decelerates the trailing part of the landslide.
The reduction in velocity begins at the rear and
progressively shifts to the front.
It is seen that the calculated sliding process
extended from the ruptured scrap in the source
zone to the deposition fan on the riverbed and
near the school, is consistent with the field
observation (Sassa et al., 2005). Clearly, there
are four major stages in the run-out process,
namely, triggering (at t 12.5 sec), accelerating

0.36

Bp

0.34
0.32
0.3
0

50

100

150
200
250
Horizontal distance : m

300

350

400

Fig 11. Distributions of the breakage potential Bp


along the sliding surface, at t = 5 sec (black line), t =
12.5 sec (blue line), t = 19 sec (golden line), t = 22.5
sec (gray line), t = 26 sec (green line), and t = 30 sec
(red line)

To get an insight into the mechanics behind


this disastrous response, Fig 11 plots the
evolution of particle breakage potential Bp.

39

in which st is the dynamic coefficient that


determines the magnitude of the gravitational
lateral soil pressure imposed to the retaining
structure. Bassanou (2000) studied the problem
experimentally and measured values of st
between 1 and 3 with smaller ones
corresponding to larger aspect ratios of the
retaining structure. hcr is the pile length covered
by the sliding soil mass, given by

Notice that Bp approaches a steady state value


of 0.30 at t > 15 seconds; this is larger than the
initial value of Bpl (computed to be 0.27 in
drained loading conditions), reflecting the
influence of the developed excess porewater
pressures. The slightly increasing breakage
potential at t > 15 seconds reveals that the
grain crushing process has been practically
terminated. The effective normal stress is not
adequate for further breakage. However, the
landslide is still accelerating due to the action of
gravity.

hcr = min(L, h )

in which L is the abovebedrock length of the


retaining structure.
Implementing equations (19) and (20) into a
beam on Winkler foundation model (Fig. 12)
and assuming that the beam axis is normal to
the basal surface, dynamic equilibrium of the
retaining wall gives the following equation

LANDSLIDEPILED WALL DYNAMIC


INTERACTION

The impact force exerted by a debris flow


against a retaining structure fixed at the
bedrock depends on the flow velocity, as well
as on the density of the flowing soil mass.
According to Armanini and Scotton (1993), the
impact force per unitary area of the structure
can be expressed as the sum of a viscous
(drag) component pd and a component due to
compressionelongation of the flowing mass pg

p = pd + p g

(27)

st
2M
+ ( st + C M ) Ast
+ pd p g = 0
2
t
z
(28)
st(z,t)

(24)

CM st(z,t)

For the computation of the first component,


several approaches have been developed
based on shallow water theory. Morison et al.
(1950) proposed the following formula for nonbreaking waves

stg(h-z)

1
p d = C d ( st ) 2 sgn( st )
2

(x,t)

(25)
x

0.5Cd

in which st is the velocity of the structural


member in the direction of flow, and Cd is the
drag coefficient depending on the Reynolds Re
and Froude Fr numbers of the flow, the
roughness, shape, and orientation of the object
with respect to flow direction. Representative
values of Cd for a cylindrical object (e.g. pile)
and an infinitely long flat plate oriented normal
to flow (e.g. retaining wall), are 1.2 and 2,
respectively.
The second component of the impact force
pg is given by

p g = st g (hcr z )

Plastic hinge

Rigid
bedrock

Fig 12: Schematic illustration of dynamic non-linear


interaction between a sliding soil mass and a
retaining piled wall. The piles are partially embedded
into the rigid bedrock.

in which Ast and st are the cross-sectional area


and mass density of the beam, respectively. CM
is the inertia coefficient representing the inertia
forces via an equivalent mass distribution along
the height of the retaining wall. CM varies
parabolically with depth obtaining its maximum
value at the top of the wall and vanishing to

(26)

40

zero at the interface with the bedrock. For a


cylindrical retaining structure (e.g. pile) with a
slenderness ratio L / d > 5 (d is the pile
diameter), the average value of CM over the
walls height is in the range of 0.81, with larger
values corresponding to larger slenderness
ratios.
In Eqn (28) M is the structural bending
moment which is assumed to be a nonlinear
hysteretic function of its curvature according to
the following equation (Gerolymos and
Gazetas, 2005).

M = st EI st + (1 st ) M y st

In the above equation, y is the value of wall


curvature at the initiation of yielding, and bst, nst
are dimensionless quantities that control the
shape of the hysteretic bending moment
curvature loop. Evidently, Eqn (30) is of the
same form as Eqn (14). In the special case of
ast = 0, My is equal to the ultimate (plastic)
bending moment of the wall Mp. For details on
the
calibration
of
the
aforementioned
parameters, the reader is referred to the work of
Gerolymos and Gazetas (2005).
The first term in the right hand side of Eqn
(28) represents the reaction force of the wall
normal to flow direction. The depthintegrated
wall reaction along the portion of the wall
covered by the flowing soil mass, is calculated
as

(29)

where EIst is the initial (elastic) bending stiffness


(also called flexural rigidity), st is a parameter
controlling the post yielding bending stiffness,
My is the value of bending moment that initiates
structural yielding in the wall, and is the wall
curvature. st is the hysteretic dimensionless
parameter that controls that controls the
nonlinear structural response of the retaining
wall. The latter is governed by the following
differential equation.

d st
1
=
1 st
y
d

nst

1
q =
hcr

hcr

2M
dz
z 2

(31)

Eqn (31) is inserted into Eqn (2) to model the


dynamic interaction of the retaining wall with the
landslide:

[bst + (1 bst ) sgn(& st )] }

(30)

(h x )
+ Td Tr T f
x

q (x x
j =1

1
(C d 2 + st g hcr )
2



) = h
+

x
t

(32)

sliding surface level, and assuming the


formation of a plastic hinge at this point, gives

in which (xxj) is the Croneckers delta


function, xj is the local horizontal coordinate of
the j retaining wall, and N is the total number of
the retaining walls distributed along the sliding
surface.
In the extreme case of a rigid structure (a
structure with infinite bending rigidity) Eqn (31)
vanishes to

q =

1
1
2
3
M = C d 2 hcr + st g hcr H ( M p M )
6
4

(34)
in which H(M) is the Heaviside step function,
suggesting that the mobilized bending moment
is equal to zero when the maximum bending
moment is reached. In reality, the maximum
bending moment is retained until to a specific
value of curvature, determined by the ductility
capacity of the wall (Priestley et al, 1996),
which in turn depends on the detailing of the
wall reinforcement. To determine the maximum

(33)

which was obtained from Eqns (25) and (26) by


neglecting the inertial force and setting st = 0.
Dynamic moment equilibrium with respect to the

41

wall covered by the landslide mass, Eqn (34) is


reformulated by setting M = Mp.

impact velocity the wall can sustain as a


function of both the plastic bending moment
and the abovebedrock length of the retaining

4M p
2 st

g hcr

2
3 Cd
C d hcr
max imp =

p g = st g (hcr z ) H ( z )

Ultimate Bending Moment Mp : MNm / m

808
7

(37)

269.4
133.3
65.7

32.1

15.4

7.2

606

3.1

404

1.0

202

failure

0
0

10

1
st g hcr3
6

15

20

25

(36)

In the right hand side of Eqn (36), the term


multiplied by the Heaviside step function H(-z)
is the lateral soil reaction acting at each point of
the embedded portion of the wall. It consists of
a viscous (first sub-term) and a hysteretic
(second sub-term) component, respectively.
Notice that negative values of z correspond to
points below the sliding surface. C is the
damping coefficient accounting for radiation of
waves emanating from the wall periphery and
propagating
through
the
soil
medium
towardsinfinity. Detailed information on the
calibration of C for a pile or an arbitrarily
shaped rigid caisson is provided by Gazetas
and Dobry (1984), and Gazetas and Tassoulas
(1987). py is the ultimate lateral soil reaction,
and s is an hysteretic dimensionless quantity
that determines the nonlinear response of the
soil, and is of the same form as Eqns (14) and
(30). Definition of this parameter is given in
Gerolymos and Gazetas (2005; 2006). For an
infinitely long retaining wall (plain strain
conditions) supporting a c soil, Rankines
theory suggests the following expression of py

Contours of max : m / s

Mp <

(35)

1
C d ( st ) 2 sgn( st ) H ( z ) + ( C st + p y s ) H ( z )
2

and

1000

1
st g hcr3
6

In the case of a compliant bedrock, Eqns


(25) and (26) for the impact force are
reformulated according to

Contours of the maximum impact velocity


against a given set of model parameters, as a
function of the local landslide thickness and
bending moment capacity of the wall, is plotted
in Fig 13.

pd =

, Mp

30

Landslide cross-sectional depth h : m

Fig 13: Contours of the maximum impact velocity


exerted on a wall founded in a rigid bedrock, as a
function of the local landslide thickness and bending
moment capacity of the wall, predicted from the
developed limit equilibrium method. The parameters
used [Eqn (35)], are: = 2 ton / m3, Cd = 1, and
st =1.

p y = 2 c tan(45 +

) + g z tan 2 (45 +

) (38)

in which c is the cohesion of the supporting soil.

42

velocity of the retaining structure are not zero.


Even though, assuming that the contribution of
the inertial force to the response of the wall is
insignificant and that the structural velocity st is
very small compared to the crosssectional
velocity of the landslide, the maximum impact
velocity can be computed by applying Broms
(1964) theory for ultimate lateral capacity. Even
if developed for piles, the applicability of Broms
method also holds for infinitely long retaining
walls. According to Broms, the lateral capacity
is reached when the moment from soil reactions
balances the ultimate bending moment at the
depth where the maximum soil resistance
develops (hinge point). The fundamental
assumption underlying the method is that the
structural movements are sufficient to fully
mobilize the plastic capacities everywhere, and
that the elastic deformations are ignored.
Applying Broms theory to our problem and
considering force and moment equilibrium with
respect to the hinge point, the following
expressions for the maximum impact velocity
imp and associated depth f of the hinge point
are obtained.

st(z,t)

CM st(z,t)

stg(h-z)
z

(x,t)
x

0.5Cd

Plastic hinge

py s

Flexible
bedrock

Fig 14: Schematic illustration of dynamic non-linear


interaction between a sliding soil mass and a
retaining piled wall. The piles are partially embedded
into the flexible bedrock.

When flexibility of bedrock is considered


(Fig. 14), Eqn (35) is no longer valid, as the
wall oscillates even if rigid. Inertial force and

Cd hcr

maximp =

p y g st hcr3 + p y hcr2 + 8 M p hcr ( py + g st hcr )


3

for
M p Mcr
M p < Mcr

for

factor larger than unity that accounts for soil


compliance. In other words, Mcr increases as
the ultimate lateral soil reaction py decreases,
while the maximum sustainable impact velocity
decreases. Eqns (40) and (41) assume that the
length of wall embedment Lemb is large enough
(Lemb > f ) to accommodate the formation of a
plastic hinge.
Fig. 15 plots the maximum impact velocity
and associated depth to fixity point for a given
set of model parameters, as a function of the
local landslide thickness and wall bending
moment capacity. The compliance of the
foundation soil corresponds to that of a very stiff
clay of undrained shear strength cu = 200 kPa.
Comparison between Figs 13 and 15 shows
that the flexibility of the foundation soil alters

and

f =

hcr
2

8M p
2 g st hcr
1+
+
2

3 py
p y hcr

1 (40)

In Eqn (39) Mcr is the minimum plastic bending


moment required for stability of the wall, given
by

M cr =

1
3 hcr
g st hcr3 1 +
g st
6
4 py

(39)

(41)

which is equal to that corresponding to a rigid


bedrock, given in Eqn (35), multiplied by a

43

discontinuity
arises
from
the
sudden
appearance of the retaining wall as is indicated
by the Croneckers delta in Eqn (32). To
overcome this problem, the above numerical
technique was used in conjunction with a
Kurganov
and
Tadmor
shockcapturing
scheme (Kurganov and Tadmor, 200) which
uses MUSCL reconstruction (Van Leer, 1979),
and the superbee algorithm (Roe, 1986) as a
TVD (Total Variation Diminishing) limiter
(Harten, 1983). It is also stressed that the
second term in the righthand side of Eqn (7)
provides physical viscocity to the problem that
damps the oscillations originating from
discontinuities and / or abrupt variations of the
basal topography, and thus contributes
beneficially to the numerical stability of the
solution.

dramatically the maximum impact velocity the


wall can sustain. Even for a relatively very stiff
soil, the maximum impact velocity could be less
than half of that corresponding to the rigid
foundation soil.

Ultimate Bending Moment Mp : MNm / m

Contours of max : m / s, and fixity depth f : m

100
22

19.8

258.6

80

139

17.6
20.9

74.5

10.8
13.2

39.7

60

15.4

5.4

= 2.44 m/s
11

40

8.8
6.6

20
0

failure

4.4
f = 2.2 m

10

15

NUMERICAL EXAMPLES
20

25

30

The capability of the model is investigated


through a number of numerical examples. The
problem under consideration is illustrated in Fig
16. The landslide material lies at rest on top of
a parabolically shaped sliding surface. The
length of the projection of the sliding surface on
the horizontal plane is 120 m, with local
inclination angles of = 31.6o at X = 0 m, and
= 0.21o at X = 120 m, with respect to the global
coordination system, and an average inclination
angle of avg = 6.82o. The surface of the
landslide material is a parabola (in section), has
a length of 30 m and a maximum thickness of 5
m at its centre (at X = 15 m). Initiation of the
landslide occurs when the basal friction
coefficient drops deliberately to = 0.05 (e.g.
due to material softening, corresponding to a
friction angle of 2.86o, that is much smaller than
the average inclination angle avg of the basal
surface), and the released soil material begins
to move down the inclined plane. Two cases
are analysed:
(a) the sliding process is free from the presence
of any retaining structure, and deposition
termination of the landslide occurs naturally.
(b) A retaining wall consisting of a row of piles
of d = 1 m in diameter and a centerto
center spacing of s = 1.5 m, with an above
bedrock height of L = 7 m and bending
moment capacity of Mp = 1 MNm, is
constructed at X = 60 m to decelerate the
sliding process. Flow of the soil mass
between the piles is allowed.

Landslide cross-sectional depth h : m

Fig 15: Contours of the maximum impact velocity


exerted on a wall founded in a compliant bedrock
and associated depth to fixity, as a function of the
local landslide thickness and bending moment
capacity of the wall, predicted from the developed
limit equilibrium method. The parameters used [Eqns
(39), (40)], are: = 2 ton / m3, Cd = 1 , st =1, and py
= 400 kPa.

The impact velocity given by Eqns (35) and


(39) should not be confused with that resulting
from collision of an individual boulder. The
collisional force of a single boulder can cause
severe local damage to the retaining structure,
and a cushion layer consisting of gravels or old
tires should be placed at its back for protection.
NUMERICAL FORMULATION

An explicit central finite difference technique


is used for the solution of Eqns (1) and (32) for
landslide motion, which are coupled with the
constitutive Eqns (14) for shear band
behaviour, and (28), (30) for non-linear bending
behaviour of the wall. The aforementioned
numerical scheme while providing accuracy of
secondorder, introduces spurious oscillations
into the solution where discontinuities or shocks
are present leading to ficticiously large
gradients of both the landslide thickness h and
the velocity . In our problem, such a

44

15

15

12

t=0s

12

15 0

20

40

60

80

100

120

15 0

t = 1.88 s

12

40

60

80

100

120

100

120

100

120

t = 1.88 s

15 0

20

40

60

80

100

120

15 0

t = 3.75 s

12

20

40

60

80

t = 3.75 s

12

15 0

20

40

60

12

80

100

120

Elevation : m

Elevation : m

20

12

t = 11.25 s

9
6
3
0

15

t=0s

15 0

20

40

60

12

80

t = 11.25 s

9
6
3
0

20

40

60

12

80

100

120

20

40

60

12

t = 16.88 s

15 0

100

120

t = 16.88 s

80

0
15 0

20

40

60

80

100

120

15 0

20

40

60

80

100

120

12

12
6

15 0

20

40

60

80

100

t = 22.50 s

t = 22.50 s

120

15

20

40

60

80

100

120

100

120

12

12

t = 30 s

t = 30 s

9
6

3
0

0
0

20

40

60

80

100

120

20

40

60

80

Horizontal distance : m

Horizontal distance : m

Fig 16: Snapshots of the landslide evolution for subcase a1 (cohesive soil: aspect ratio = 0.2).

Fig 17: Snapshots of the landslide evolution for subcase a2 (granular soil: aspect ratio = 0.8).

45

15

15
12

12

t=0s

9
6

wall

15 0

20

40

60

80

100

15 0

120

t = 1.88 s

12

20

40

60

80

100

120

15

t = 3.75 s

12

60

80

100

120

100

120

100

120

t = 1.88 s

20

40

60

80

t = 3.75 s

12

0
20

40

60

12

80

100

120

Elevation : m

15 0

Elevation : m

40

t = 11.25 s

9
6
3
0

15 0

20

40

60

12

80

t = 11.25 s

9
6
3
0

20

40

60

12

80

100

120

15 0

20

40

60

12

t = 16.88 s

80

100

120

t = 16.88 s

0
15 0

20

40

60

80

100

120

15 0

20

40

60

80

100

120

12

12

t = 22.50 s

t = 22.50 s

0
15

20

12

15

wall

15

t=0s

20

40

60

80

100

120

15 0

20

40

60

80

100

120

100

120

12

12

t = 30 s

t = 30 s

0
0

20

40

60

80

100

120

20

40

60

80

Horizontal distance : m

Horizontal distance : m

Fig 19: Snapshots of the landslide evolution for subcase b2 (granular soil: aspect ratio = 0.8). A row of
piles are located at X = 60 m.

Fig 18: Snapshots of the landslide evolution for subcase b1 (cohesive soil: aspect ratio = 0.2). A row of
piles are located at X = 60 m.

46

case a1, the flow remains expanding from its


initiation to its termination. The length of the
deposition zone amounts to 68 m, that is 1.5
times larger than in case a1 (cohesive soil),
covering the region between X = 38 m and X =
106 m. The resulting runout distance is now
76 m, that is 15 m larger than that of case a1.
The role of the retaining wall on the breaking
of the landslide movement (Figs 18 and 19) is
double. The existence of a piledwall at X = 60
m reduces both the runout distance of the
landslide and the extent of its deposition zone.
However, the contribution of the wall to the
termination of the landslide is significantly more
profound in case a2 (cohesive soil) than in case
b2 (granular soil). The runout distance and
associated length of the deposition zone
amount to 49 m and 36 m for case a2 and to 69
m and 64 m for case b2, respectively.

in both cases the bedrock is assumed to be


rigid, while in the second case the retaining wall
is assumed to behave as a rigidperfectly
plastic material. This means that q is
calculated directly from Eqn (33), and thus
differential Eqns (28) and (30) are simply
eliminated from the model. As the actual
problem is 2D in nature while the proposed
model is 1D, averaging the effect in 2D over the
distance between the piles is necessary. To this
end, Eqn (33) for the depthintegrated wall
reaction is modified as

q eq =

d
q
s

(42)

Eqn (42) suggests that the row of piles is


represented by an equivalent massive wall, the
width of which is equal to the pile diameter
multiplied by the ratio d / s. The actual force
and moments on the piles are then calculated
by dividing the computed ones by the
aforementioned ratio.
The values of the parameters used in the
analyses are: = 2 ton / m3, s = 0.5, = 0.2, d
= 5, = 0.08, st = 1.5, Cd = 1.5, b = 0.5, n = 3,
and uy = 10-3 m. Four sub-cases are
investigated regarding the effect of the aspect
ratio on the landslide evolution. Sub-cases a1
and b1 for = 0.2, representative for cohesive
soils, and subcases a2 and b2 for = 0.8,
representative for granular soils. The entire
computational area is divided into 120 cells,
resulting in a cell length of 1 m, while the time
step is taken equal to a very small value, t =
3.75 10-4 s, to ensure numerical stability.
The results of the analyses are plotted in
Figs 16 to 20. Fig 16 illustrates the simulated
sliding process for case a1 (cohesive soil: =
0.2). After its termination at t = 30 sec, the
landslide has a length of 45 m, that is 1.5 times
larger than its initial length of the sliding mass,
covering the region between X = 46 m and X =
91 m. Observe that at t < 17 sec the flow is
contracting, as the avalanche body has not yet
reached the gently sloping part of the sliding
surface. After a runout distance of 61 m, the
basal friction is large enough to bring the
avalanche to rest.
The corresponding results for case b1
(granular soil: = 0.8) are presented in Fig 17.
As the rheological component of the sliding soil
mass is now more intense, compared to that of

Impact moment at the pile base : kNmo

500
400
300
200
100
0
0

10

15

20

25

30

time : sec

Fig 20: Time histories of the impact moment


developed at the pile base for cases b1 (cohesive
soil: aspect ratio = 0.8) with black line, and b2
(granular soil: aspect ratio = 0.8) with grey line.

Fig 20 plots the time histories of the impact


moment developed at the pile base during the
sliding process, for cases a2 and b2. The figure
reveals that the frontal edge of the landslide
reaches the wall at t = 8.8 sec in case a2, later
than in case b2 where the wall is hit at t = 7.5
sec. The maximum impact moment is computed
to 431 kNm for case a2, and to 132 kNm for
case b2. Both values are significantly lower
than the capacity of the wall. The considerably
smaller bending moment in case b2 compared
to that in case a2, is attributed to the expanding
nature of the flowing process, which results in

47

significantly compared to that for a rigid


bedrock.
The capability of the model is demonstrated
through a set of numerical examples,
highlighting the role of a retaining wall on the
landslide evolution. It is shown that the
presence of the wall results in the reduction of
both the runout distance and the deposition
length of the landslide. The reduction becomes
more profound with increasing cohesiveness of
the landslide material. As the actual problem
studied in this paper is 2D in nature, validation
of the proposed 1D model through experiments
is strongly recommended.

low landslide thicknesses and thus in a small


impact force on the wall.
CONCLUSION

In this paper we developed a model for


seismic triggering, evolution and stopping of a
landslide with a retaining structure. The
evolution of the landslide is modeled via an
extended SavageHutter model coupled with
the MohrCoulomb sliding law for the frictional
deformation of the material within the shear
band.
The capability of the model is investigated
through prediction of the HigashiTakezawa
landslide. Three scenarios are analysed
regarding the location of the sliding surface and
the susceptibility of sand to grain crushing: (a)
shear band within the sand layer, but sand not
susceptible to grain crushing, (b) shear band
within an assumed thin silt layer at the top of
the siltstone, but sand not susceptible to grain
crushing, and (c) shear band within the sand
layer, but sand susceptible to grain crushing.
The residual displacement is calculated to be
0.65 m and 3.4 m for the first and second
scenario, respectively. The observed 100 m
displacement of the landslide, associated to a
shear velocity of 15 m / sec (20 sec after the
triggering), is only reproduced with the third
scenario, despite the residual friction angle of
the silt [scenario (b)] being 13 degrees smaller
than that of sand.
The developed depthintegrated model for
landslide kinematics is then appropriately
modified to implement the dynamic interaction
between a rapidly deformed landslide and a
retaining piledwall. The derived system of
differential equations is solved numerically
utilizing an explicit central finite difference
technique in conjunction with a shockcapturing
scheme to avoid spurious oscillations
originating from abrupt variations of the model
properties.
A limiting equilibrium approach to the
problem is also presented, resulting in design
diagrams of the maximum impact velocity the
wall experiences as a function of the local
landslide thickness and structural strength of
the wall. It is shown that bedrock compliance
influences greatly the maximum impact velocity.
Even for a relatively stiff bedrock, the maximum
impact velocity the wall can undertake reduces

ACKNOWLEDGMENTS

This paper is a partial result of the Project


PYTHAGORAS I / EPEAEK II (Operational
Programme for Educational and Vocational
Training II) [Title of the individual program:
Mathematical and experimental modeling of the
generation,
evolution
and
termination
mechanisms of catastrophic landslides].This
Project is co-funded by the European Social
Fund (75%) of the European Union and by
National Resources (25%) of the Greek Ministry
of Education.
REFERENCES
Armanini A, Scotton P (1993) On the dynamic
impact of a debris flow on structures. Proc. Of
the XXV IAHP Congress, Tokyo, 203210.
Baldwin J E, Donley H F, Howard T R (1987) On
debris flow/avalanche mitigation and control,
San Francisco Bay area, California. Reviews in
Engineering Geology. Geol. Soc. Of America,
VII: 223236.
Bassanou M (200) Landslides dynamics
Mathematical and experimental simulation of
debris flow. Ph. D. Thesis, National Technical
University, Athens, Greece (In Greeks).
Broms B B (1964) Lateral resistance of piles in
cohesive soils. J. soil mech. Fdns. Div., ASCE,
90: 2763.
Chen H, Lee C F (2002) Runout Analysis of Slurry
Flows with Bingham Model, J. of Geotechn. And
Geoenv. Eng., ASCE, Vol. 128(12): 10321042.
Fraccarollo L (1995) Un modello matematico
bidimensionale per la mappatura del riscio da
colata di detriti. GNDCI, Linea 1: 97105 (in
Italian).

48

Jan C D (1997) A study on the numerical modeling


of debris flow. Debris Flow Hazards Mitigation:
Mech., Pred. and Assessment, ASCE.

Gazetas G. and Dobry R (1984) Simple radiation


damping model for piles and footings. Journal of
Engineering Mechanics, ASCE, Vol. 110:
937956.

Jin M, Fread D L (1997) 1D routing of mud/debris


flow using NWS FLDWAV model. Debris Flow
Hazard Mitigation: Mech.,Pred. and Assessment,
ASCE.

Gazetas G, Tassoulas J L (1987) Horizontal


damping of arbitrarily shaped embedded
foundations.
Journal
of
Geotechnical
Engineering, ASCE, Vol. 113(5): 458475.

Kokusho T, Ishizawa T (2005) Energy approach to


slope failures and a case study during 2004
Niigataken Chuetsu earthquake. Proceedings
of
Geotechnical
Earthquake
Engineering
Satellite Conference Osaka, Japan, 10
September 2005, Performance Based Design in
Earthquake Geotechnical Engineering: Concepts
and Research, ISSMGE: 255-262.

Gerolymos N, Gazetas G (2005) Constitutive Model


for 1-D Cyclic Soil Behavior Applied to Seismic
Analysis of Layered Deposits. Soils and
Foundations 45(3): 147159.
Gerolymos N and Gazetas G (2007) A Model for
Grain Crushing Induced Landslides Application
to Nikawa, Kobe 1995. Soil Dynamics and
Earthquake
Engineering
(accepted
for
publication).

Kurganov A, Tadmor E (200) New highresolution


central schemes for nonlinear conservation laws
and convectiondiffusion equations. J. Comp.
Phys. 160: 214282.

Gerolymos N, Vardoulakis I and Gazetas G. (2007)


A thermo-poro-viscoplastic shear band model
for triggering and evolution of catastrophic
landslides. Soils and Foundations 47(1).

Laigle D, Coussot P (1997) Numerical modelling of


mudflows. J. of Hyd. Eng.

Gray J.M.N.T., Wieland M., Hutter K. (1999) Gravity


driven free surface flow of granular avalanches
over complex topography, Proc. R. Soc. London
A 455, 1841-1874.

Morison J R, OBrien M P, Johnson J W, Schaaf S A


(1950) The force exerted by surface waves on
piles. Petroleum Transaction, American Institute
of Mining Engineers 189: 149154.

Hadush S, Yashima A, Uzuoka R (2000)


Importance of viscous fluid characteristics in
liquefaction induced lateral spreading analysis.
Computers and Geotechnics 27: 199224.

OBrien J S, Julien P J, Fullerton W T (1993) Two


dimensional
water
flood
and
mudflow
simulation. J. of Hyd. Eng., Vol. 119(2).
Priestley M J N, Seible F, Calvi G M (1996) Seismic
design and retrofit of bridges. John Wiley and
Sons, Inc., New York

Harten H (1983) High resolution schemes for


hyperbolic conservative laws. J. Comp. Phys.
49: 357385.

Sassa K, Fukuoka H, Wang F, Wang G (2005)


Dynamic properties of earthquake-induced
large-scale rapid landslides within past landslide
masses. Landslides 2: 125-134.

Hungr O., Morgan G C, Van Dine D F and Lister D R


(1987) Debris flow defences in British
Columbia. Reviews in Engineering Geology
Geol. Soc. Of America, VII: 201222.

Pudasaini S. P., Hutter K. (2003) Rapid shear flows


of dry granular masses down curved and twisted
channels, J. Fluid Mech., 495, 193-208.

Hutter K, Greve R (1993) Twodimensional


similarity solutions for finitemass granular
avalanches with Coulomb and viscoustype
frictional resistance. J. Glac. 39: 357372.

Roe P L (1986) Characteristic based schemes for


theEuler equations. Ann. Rev. Fluid Mech. 18:
p337.

Hutter K, Koch T (1991) Motion of granular


avalanche in an exponentially curved chute:
experiments and theoretical predictions. Phil.
Trans. R. Soc. London, A 334: 93138.

Savage S B, Hutter K (1989) The motion of a finite


mass of granular material down a rough incline.
J. Fluid Mech. 199: 177215.

Iverson R M (1997) The physics of debris flows.


Rev. Geophys. 35: 245296.

Savage S B, Hutter K (1991) The dynamics of


avalanches of granular materials from initiation
to runout. Part I: Analysis, Acta Mechanica 86:
201223.

Iverson R. M., Delinger R. P. (2001) Flow of


variably fluidized granular masses across threedimensional terrain. 1. Coulomb mixture theory,
J. Geophys. Res. B, 106, 537-552.

Toro E F (1999) Riemann solvers and numerical


methods for fluid dynamics. SpringerVerlag,
Berlin Heidelberg.

49

Van Leer B (1979) Towards the ultimate


conservative difference scheme. A second order
sequel to Godunovs method.J. Comp. Phys.
32: 101136.

Tsukamoto Y, Ishihara K (2005) Residual strength


of soils involved in earthquakeinduced
landslides. Proceedings of Geotechnical
Earthquake Engineering Satellite Conference
Osaka,
Japan,
10
September
2005,
Performance Based Design in Earthquake
Geotechnical Engineering: Concepts and
Research, ISSMGE: 117-123.

Zwinger
T
(2000)
Dynamik
einer
Trockenschneelawine auf beliebig geformten
Berghngen. Ph. D. Thesis, Vienna University
of Technology (in German).

Uzuoka R, Yashima A, Kawakami T, Konrad J M


(1998) Fluid dynamics based prediction of
liquefaction
induced
lateral
spreading.
Computers and Geotechnics 22: 243282.

50

Interaction of Earthquake-Triggered Landslide with Foundation-Structure


Systems
R. Kourkoulis, F. Gelagoti, I. Anastassopoulos, G. Gazetas
National Technical University of Athens, Greece

Abstract
The paper studies the effects of earthquake induced landslides on structures founded on
the vicinity of slope crests. Plane-strain dynamic analyses are performed utilizing fully non
linear finite elements. The model is calibrated against published data to simulate the
observed strain-softening behavior of soil during a seismic event and under the action of
gravitational forces. The foundation is modeled as a flexural beam and the possibility of
sliding between the foundation and the underlying soil is considered through the use of
special gap elements. The influence of foundation type (shallow or piled), on the position of
the failure surface and the produced soil-displacements is explored parametrically. The
analysis shows that the use of mat foundation compared to that of isolated footings is
generally more suitable. Properly designed piled foundations can also significantly enhance
the seismic response of the structure.

INTRODUCTION

known that the resistance of soils whose


strength is characterized by different peak and
residual values, is progressively reducing with
increasing strain (Terzaghi and Peck, (1948),
Skempton (1964), Bjerrum (1967). The complex
mechanism of progressive failure of slopes
apparently cannot be modelled with simplified
limit equilibrium techniques and rather
necessitates advanced numerical methods
capable
to
describe
strain-localization
phenomena.
To simulate the progressive soil-failure and
the shear-zone development a popular
methodology is the use of Finite Elements
(Hoeg 1972, Lo and Lee 1973, Dounias et al
1988, Chen et al 1992, Modaressi et al 1995
,Potts et al 1997, Loukidis et al 2003,
Troncone, 2005 Pradel et al 2005).All the above
studies were performed under static conditions
and only the elements in the vicinity of the
expected failure surface were modelled so as to
obey a certain strain-softening law while the
rest of the soil model was assumed to behave
elastically.
The present study, utilizes a fully non-linear
finite element model with a strain-softening

The seismic bearing capacity of shallow


footings on the vicinity of sloping ground has
been thoroughly examined. Sarma and Chen,
(1995), Sarma (1999), Askari and Farzaneh,
(2003), Kumar and Rao (2003)), by means of a
pseudostatic limit equilibrium approach that
takes into account the inertia of soil mass,
concluded that the bearing capacity is minimal
when the footing is located at the edge of the
slope, and increases as the footing is carried
away from it until it ultimately converges to its
level-ground value.
Despite its broad acceptance, limit
equilibrium technique is inappropriate to
effectively simulate the soil-structure-interaction
effects that take place during earthquakes. First
of all it is inherently incapable to capture the
actual strain-development phenomena. The
method usually requires a pre-assumed sliding
surface that is not the natural product of
strength reduction due to strain accumulation.
Moreover the pseudostatic approach is not
applicable to soils that soften with increasing
number of cycles (Loukidis et. al (2003)), such
as those examined in the present study. It is

51

surface, while at the same time the


computational
time
was
exponentially
increased.

material law. The potential separation of the


foundation from the soil, due either to uplift or to
the downward slope movement during
landslide, is also taken into account. Since this
model can exhibit non-uniform straining, it
realistically captures the mechanism of
progressive slope failure and the effects of the
foundation on the position of the generated
failure surface. Because the sliding surface is
not pre-defined but is product of strain
softening, the model encapsulates both the
effect of the structure on the produced
displacement-field and on the position of the
failure surface itself. The soil is at this stage
assumed to be dry, neglecting the effect of pore
pressure build up due to cyclic loading.

SOIL MODELING AND CALIBRATION

A non-associated flow rule is adopted for the


soil. The pre-failure behavior and the loadingunloading of the soil obey the theory of
elasticity, while strain softening and the postfailure behavior are modeled with s MohrCoulmb failure criterion. The model parameters
are the cohesion, c, the friction angle and the
dilatancy angle . Strain softening is
incorporated into the finite element code
through a user defined subroutine that reduces
the strength parameters c and with increasing
plastic strain (Figure 1a). The material behavior
is calibrated against a viscous-plastic model for
the calculation of the strain and strength of
strain-softening cohesive soils developed by
Gerolymos et al., (2007), Based on several
experimental results (e.g. Lupini et al.,1981,
Bishop et al., 1971, Bromhead and Curtis,
1983, Skempton, 1985, Tika and Hutchinson,
1999) and utilizing an artificial neural network,
they proposed analytical relationships for the
calculation of the strength parameters of
cohesive
soils
depending
on
their
characteristics. The residual friction angle (Fig.
1b) has been found to be a function of the clay
content (CF ) and the clay activity (A):

FINITE ELEMENT MODEL

The analyses are performed in the finite


element code ABAQUS (2001) Plane strain
dynamic analyses are executed, utilizing a
fully non-linear elasto-plastic model. A 30 m
high, 23o steep slope is analysed. Two
different soil strata are considered. The top 40
m layer exhibits strain softening behaviour
according to the material law described below.
The bottom stratum is considered to be
elastic, while its thickness till the bedrock is 20
m
Quadrilateral 4-noded plane strain elements
are used for the representation of the soil while
the foundation slab is modeled with beam
elements. The interface between the foundation
slab and the underlying soil is modeled with
frictional elements which allow both slippage
and separation between the soil and the
foundation nodes. Both shallow and piled
foundations are examined. The piles are
modeled as beam elements; their diameter is
0.60 m, while their length extends to 30 meters.
This way the piles are founded within the stiff
soil layer below the level of the observed slope
failure surface. Free field conditions are applied
at the model boundaries. The optimum element
dimension that would ensure computational
efficiency without endangering the accuracy of
the simulation was parametrically investigated.
A size of L=0.5 m has been finally selected. It
was shown that the use of smaller elements
had no effect on the position of the failure

= P / C

(1)

where Ip the plasticity index.


Skempton (1985) suggests typical values of
displacements during various stages of the
ring shear test. The critical state friction angle
is calculated according to Mitchell (1976) as:
cv arcsin [ 0.6 0.14log(IP 5)]

(2)

Finally, the peak value of the friction angle


was found to be best calculated as follows:
tan P

tan cv

52

= 1 0.85 ln

2 ,
OCR

OCR > 2

(3)

PARAMETRIC ANALYSES

Infinite Strip

Scope of this study is a parametric


investigation of the effect of foundation type on
the failure surface and the generated soil
displacements for different levels of ground
shaking. Three types of structures are
investigated:
1) infinite strip of width 20 m and load q = 20,
40 and 80 kPa per unit area. The strip is
located at distances s=5 m, 8 m and 11 m from
the crest. The purpose of this study is to
examine whether and to what extend the load
and the location of the foundation affect the
position of the generated shear zone and the
resulting soil-displacement pattern.
2) a single storey structure of width 20 m and
load q=20 kPa per unit area founded on (a)
isolated footings and (b) rigid foundation slab.
Here the effect of foundation type is
investigated both in terms of stresses and
displacements on the structure.
3) A slab of width 20 m and load q=20 kPa per
unit area supported through piles. This set of
analyses aims to assess whether piled
foundation is a method of enhancing the
stability of structures in the vicinity of sliding
slopes.
The JMA accelerogram recorded during the
devastating Kobe 1995 earthquake (Fig 2)
scaled at PGAs of 0.5 g and 0.8 g is used as
the excitation motion of the model.

Position of the failure surface :


When a 20 m width rigid footing of q = 40 kPa
lies at 5 m distance from the slope crest the
picture of the generated plastic deformations is
notably changed. Fig 5a illustrates the new
failure surface when the slope is excited by
JMA accelerogram (PGA 0.8g). It is clear that
the footing causes the failure surface to deviate
from its free field position, as it imposes an
extra load on the slope crest which in turn
provokes failure of its underlying soil. As the
footing is taken further away from the crest at a
distance of 8 m (Fig. 5b), the failure surface
seems to be less affected until it finally returns
to its initial free field position once the footing
has been moved far away ( Fig. 5c). At that
distance, the influence of the footing on the
slope failure is minimal.
When the footing load is reduced to the half
(q = 20 kPa) the failure surface (Figure 6a) is
practically unaffected by the footings presence.
Conversely, in case the foundation load is large
(80 kPa), the shear zone is indeed influenced
by the footing: now the shape of the generated
failure surface is reminiscent of a bearing
capacitytype failure (Fig. 6c). As can be
deduced by the intensity of plastic deformations
generated underneath the footing (Figs 5 and
6), the bearing capacity of the ground increases
as the footing is moved away from the crest.

RESULTS

Footing Response :
Fig 7 plots the horizontal sliding of the soil
nodes underneath the two footing edges (nodes
1 and 2) along with the sliding of the footing
itself (Node 3). As expected, soil node 1 which
lies on the side of the sloping ground
systematically displaces the most, while soil
node 2 displaces less as it is less affected by
the slope movement. In case of the immense
JMA input motion scaled at 0.8 g., the
displacement of node 1 keeps increasing with
time an evidence of landslide. On the other
hand, as the footing detaches from the ground
once the friction force is exceeded, the footing
node 3 displaces less than the underlying soil
node 1. Note also that when the excitation
motion is the JMA record scaled at 0.5g the
landslide is not triggered (after the end of the
earthquake t > 27 sec the soil-displacements

We begin with the case when no structure is


founded on the top of the slope crest (FreeField case). Fig 3a plots the generated field
displacements when the input motion is the
JMA with PGA 0.5g, while Fig 3b shows the
same distribution when the peak ground
acceleration is 0.8g. The plastic strain
distribution is depicted in Fig 4. It is evident that
when the model is subjected to the JMA0.5g no
landslide is triggered (the maximum plastic
strain doesnt exceed 4 %), while the maximum
soil-displacement is 53 cm. On the contrary
when the excitation is the JMA0.8g the
formation of the shear zone is clear (40 %
plastic strain) and the maximum field
displacement after the application of the gravity
load is 1.8 m!

53

remain constant under the action of the gravity


load)

foundation slab decreases significantly the


developed moments compared to the use of
isolated footings. Greater are the differences at
the peripheral columns where the moments at
the top with the use of isolated footings are
double those developed with the mat
foundation.

One-Storey Building
The frame is considered to be elastic. Emphasis
is given on how the different types of foundation
affect the developed stresses on the structure
and the generated soil-displacements. It is
mentioned that in some cases the developed
stresses are greater than the actual resistance
of the cross-section. However, for comparison
reasons we keep the geometric characteristics
of the cross-section constant.

Displacements on the structure:


Fig 12 and 13 present the displacement time
history at particular points on the structure. Only
results for the JMA 0.8g excitation are depicted.
It is clear that the type of foundation controls
the magnitude of the differential settlement.
Note that when the foundation is a rigid slab the
differential settlement is overall smaller than the
one observed when isolated footings are used.
Moreover with the use of mat-foundation the
structure displaces as a rigid body which
results in general smaller deformations on the
frame.

Position of the failure surface :


Fig 8 displays the field plastic-strain distribution
when the excitation motion is the JMA
accelerogram scaled at 0.8g. In the upper
Figure the structure is supported on isolated
footings while in the bottom Figure the same
structure is founded on a rigid slab. It is clear
that the position and the shape of the shear
zone doesnt change with the foundation-type.

Piled Foundation
The piled foundations are only limited
examined as an alternative foundation method.
Our scope is to detect its potential benefits
compared to the shallow foundations. The piles
are considered to behave elastic. They go up to
depth 30 m where they reach the rock stratum
and their diameter is 60 cm.
The connection between the pile-node and
the soil-node is accomplished through properly
calibrated non-linear springs that follow
published p-y curves (Matlock, 1970). When
the spring deformation exceeds a certain value
(given by the empirical p-y curve) the spring
stiffness drops to zero. With this approach the
separation between pile and soil is achieved at
high values of soil displacement. It is believed
that with this numerical trick the actual 3-D
behavior of the pile-soil system is more
realistically described. Fig 14 depicts the typical
spring behavior during a seismic event that
precedes a landslide for two characteristic
springs (one at the top and one at the bottom of
the pile).
Figure 15 compares the final field
displacements when a shallow and a piled
foundation are used. Figure 16 presents the
horizontal displacement time histories of a
structure with (a) a shallow and (b) a piled
foundation respectively. It is noted that the
presence of piles improves the soil-

Field Displacements:
Fig 9 portrays the final soil-displacements when
the model is subjected to the JMA time-history
scaled at 0.8g. It is clear that the different
foundation type doesnt affect the resulting
displacement profile. The trends are reserved
when the input motion is the JMA0.5g.
Stresses on the structure:
Figures 10 and 11 depict the envelope of
maximum-bending-moments on the columns of
the frame when the input motion is the JMA
record scaled at 0.5g and 0.8g respectively.
The following are observed:
When the input is the JMA at 0.5g (an
excitation that doesnt trigger landslide), the
differences in the developed moments for the
two foundation-types are not significant. It is
believed that since this seismic time-history
doesnt evoke significant soil displacements,
the developed stresses are essentially the
result of the inertial response of the frame.
On the other hand, when the excitation is the
JMA 0.8g the developed soil displacements are
very high. This results further kinematic stress
on the frame due to the imposed displacements
of the landsliding soil stratum. In this case the
type of foundation is crucial. The use of rigid

54

Lo, K. Y. And Lee, C.E. (1973), Stress analysis and


slope stability in strain-softening materials.
Geotechnique 23, No1, 1 11

displacement pattern and decreases the peak


values. Most importantly the pile foundation
seems to relieve the supported structure which
deforms less.

Loukidis D., Bandini P. and Salgado R. (2003),


Stability of seismically loaded slopes using limit
analysis Geotechnique 53, No. 5, 463479

REFERENCES

Lupini J.F., Skinner A.E. and Vaughan P.R.


(1981),The drained residual strength of cohesive
soils. Geotechnique, 31(2), 181213

Anastasopoulos I. (2005), Behaviour of Foundations


over Surface Fault Rupture: Analysis of Case
Histories from the Izmit (1999) Earthquake,
Proceedings of the 16th International Conference
on Soil Mechanics & Earthquake Engineering,
Osaka, Japan, September 1216, 2005, pp.
26232626.

Mitchell J. K. (1976): Fundamentals of soil


behaviour, John Wiley and Sons, New York, 422.
Modaressi H., Faccioli E., Aubry D., Noret C. (1995),
Numerical modelling approaches for the analysis
of earthquake triggered landslides. Proceedings
of the Third International Conference on Recent
Advances
in
Geotechnical
Earthquake
Engineering and Soil Dynamics, St. Louis,
Missouri, II(INVLE.03), 833843

Anastasopoulos, . (2005), Fault RuptureSoil


FoundationStructure
Interaction,
Ph.D.
Dissertation, School of Civil Engineering,
National Technical University, Athens.
Askari, F. & Farzaneh, O. (2003). Upper-bound
solution for seismic bearing capacity of shallow
foundations near slopes Geotechnique 53, No. 8,
697702

Fardis N., Georgarakos P., Gazetas G.,


Anastasopoulos I.(2003) Sliding Isolation of
Structures: Effect of horizontal and vertical
acceleration. Proceedings of the fib SymposiumMay 6-8 Athens, Greece (in cd rom)

Bjerrum L. (1967) Progressive failure in slopes of


over-consolidated plastic clays and clay shales,
J. Soil Mech. Fdn Div. ASCE 93, 349

Potts D.M., Dounias G.T., and Vaughan P.R. (1990),


Finite element analysis of progressive failure of
Carsington Dam embankment. Geotechnique 40,
No 1, 79 101

Bromhead E.N. and Curtis R.D (1983): A


comparison of alternative methods of measuring
the residual strength of London Clay. Ground
Engineering, 16.

Potts D.M., Kovacevic N. And Vaughan P.R. (1997),


Delayed collapse of cut slopes in stiff clay.
Geotechnique 47, No. 5, 953982

Budhu, M., Al. Karni,Y.(1993) The seismic bearing


capacity of soils, Geotechnique, 43 (1), p.p. 181187

Potts, D.M. and Zdravcovic L. (1999), Finite element


analysis in geotechnical engineering: theory.
London: Thomas Telford

Chen. Morgenstern N. R. and Chan D. H. (1992)


Progressive failure of the Carsington Dam: a
numerical study. Can. Geotech. J. 29, 971988

Pradel D., Smith P.M., Stewart J.P. and Raad G.


(2005), Case History of Landslide Movement
during the Nothridge Earthquake JGGE, ASCE,
11, 1360-1369

Dounias G.T., Potts D.M. and Vaughan P.R. (1988)


Finite element analysis of progressive failure: two
case studies Comput. Geotech. 6, 155175.
Gerolymos N., Vardoulakis I. and Gazetas G. (2007)
A
thermoporovisco-plastic
shear
band
model for seismic triggering and evolution of
catastrophic landslides. Soils and Foundations
47 (1).

Sarma S.K. (1999), Seismic bearing capacity of


shallow strip footings adjacent to a slope.
Earthquake Geotechnical Engineering, Seco e
Pinto (ed), Balkema, Rotterdam, 309-313. Proc.
Second International Conference on Earthquake
Geotechnical Engineering, Lisbon, Portugal

Hoeg, K. (1972). Finite element analysis of strain


softening clay, J. Soil Mech. Fdns Div. Am. Sot.
Ciu. Engrs,
43-59.

Sarma S.K. and Kourkoulis R.S (2004) Investigation


into the prediction of Sliding Block Displacements
in Seismic Analysis of Earth Dams, Proceedings
13 WCEE, Vancouver B.C, (in cd-rom)

Larsson R., Runesson K. and Sture S. (1991), Finite


element
simulation
of
localized
plastic
deformation. Arch. Aplpl. Mechanics 61, 305
317.

Skempton A.W. (1985): Residual strength of clays in


landslides, folded strata and the laboratory,
Geotechnique, 35(1).
Terzaghi K. and Peck R. B.(1948) Soil Mechanics in
Engineering Practice,. New York: Wiley

55

Tika Th.E., and Hutchinson J.N. (1999), Ring shear


tests on soil from the Vaiont landslide slip
surface. Geotechnique, 49(1)
Troncone A. (2005), Numerical analysis of a
landslide in soils with strainsoftening behavior,
Geotechnique 55, No.8, 585596
(a)

(b)

Peak strength

Mobilized friction angle

Critical state

cv

Slow residual state


(at large displacements)

r,s
OC clay

r,d
Fast residual State
(at large velocities)

10

Slow residual friction angle r,s : deg

40
35

A = 0.75 - 1.2

30

A = 1.5 - 4.3

25

A = 0 - 0.6

20
15
10
5
0

400

Displacement : mm

20

40

60

80

Clay fraction CF : %

Fig 1 : (a) Typical behavior of strain-softening soils (b) Curve for the calculation of the slow residual angle as
a function of clay fraction and parameter A

0.8
0.6

a:g

0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
0

10

15

20

t : sec

Fig 2 : JMA recorded time-history, Kobe 1995, Japan

Fig 3 : Contours of horizontal displacement for the free field case when the model is subjected to the JMA
accelerogram scaled at (a) 0.5 g and (b) 0.8g.

56

Fig 4 : Contours of plastic deformations generated for the free field case when the model is excited by the
JMA recode scaled at (a) 0.5g and (b) 0.8g.

Fig 5 : Contours of plastic deformations generated for the case of a 20 m wide footing of q=40kPa lying at a
distance from the crest of (a) 5m, (b) 8 m, (c) 11 m, subjected to 0.8g JMA excitation.

57

Fig 6 : Contours of plastic deformations generated for the case of a 20 m wide footing lying at 8m from the
crest with load (a) q=20kPa (b) q= 40 kPa, (c) q = 80 kPa, subjected to the JMA excitation at 0.8g.

3
1

(a)

(b)

u (m )

u :m 1.60

u :m 0.50
0.40

u (m )

1.20

0.30
0.80

0.20
node 1
node 2
node 3

0.10
0.00
0

10

20

node 1
node 2
node 3

0.40
0.00
0

30

10

20

30

t:s

t:s

Fig 7: Horizontal displacements at nodes 1, 2 and 3 respectively when the input motion is (a) a JMA scaled at
0.5 g and (b) JMA scaled at 0.8 g.

58

(a)

(b)

Fig 8: Contours of plastic deformations in the case of a one-storey building at 8m from the crest with load
q=20kPa subjected to the JMA 0.8g, when the foundation is (a) mat-foundation and (b) isolated
footings.

umax = 2 m !

umax = 2 m

umax = 2 m !

umax = 2 m

Fig 9: Contours of horizontal field displacements in the case of a one-storey building at 8m from the crest with
load q=20kPa subjected to the JMA 0.8g, when the foundation is (a) mat-foundation and (b) isolated
footings.

59

Fig 10: Envelope of maximum bending-moments on a structure with (a) isolated footings and (b) matfoundation when the excitation is the JMA record at 0.5g.

Fig 11: Envelope of maximum bending-moments on a structure with (a) isolated footings and (b) matfoundation when the excitation is the JMA record at 0.8g.

60

u1 (m)

Point D

Point C

Point

Point

U 1(m)
u1 : m

1.2

Point

0.8

Point
Point C

0.6

Point D

0.4

0.2

0
0

10

15

20

25

30

35

t (sec)
: sec

-0.2
0.1

0
0

10

15

20

25

30

Point C

u2 (m)
Point

Uu2
2(m)
:m

-0.1

Point D

Point

35

Point D
Point C

-0.2

-0.3

Point



-0.4

Point

t (sec)

-0.5

t : sec

Fig 12: Displacement time-histories at points A, B, C and D on a structure with isolated footings when the
excitation is the JMA record at 0.8
1.2
1.2

Point A
Point C

1.01

U (m)

u:m

0.8
0.8

Point B

0.6
0.6
0.4
0.4

0.2
0.2

0
0

10

15

10

20

15

20

-0.2

: sec
t t(sec)
0.7
0.6
0.6

()
: degrees

0.5
0.4
0.4
0.3
0.2
0.2
0.1

0
0

-0.1

10

10

15

15

20

20

25

25

30

30

35

35

t (sec)

t : sec

Fig 13: Displacement time-histories at points A, B, C and D on a structure with mat-foundation when the
excitation is the JMA record at 0.8g.

61

P (KN)

P : kN

300
300
200
200
100
100
0
0.05
0.05

0.1
0.1

0.15
0.15

-100

-100

0.2
0.2

y:m

y (m)

P (KN)

P : kN

1010
0
-0.002
-0.0015
-0.001
-0.0005
-0.002
-0.0015
-0.001 -0.0005
-10
-10

0.0005
0.0005

0.001
0.001

y:m

y (m)

-20
-20

-30
-30

Fig 14: p-y curves for two characteristic springs

()

(a)

(b)
()

Fig 15: Final soil displacements when (a) a shallow and (b) a piled foundation are used and the excitation
motion is the JMA record scaled at 0.8g.

62

(a)
U (m)

(b)
1

0.8

uu = 0.9
mm
=0.9

0.8

found

u ==0.7
0.7 mm
ufound

0.6

u:m

0.6

u:m

usoil =0.95 m

0.4

0.2

0.4

0.2

0
0

10

15

20

25

30

35

-0.2

10

15

20

25

30

35

t (sec)

-0.2

t : sec

Fig 16: Horizontal Displacement time-histories on (a) shallow and (b) piled foundation when excited by the
JMA record scaled at 0.8g.

(deg)
1

: degrees

0.8
0.6
0.4

shallow foundation

0.2

piled foundation

0
0

10

20

30
t (s)
t : sec

Fig 17: Compare the time-histories of the tilt angle between a shallow and a piled foundation when the input
motion is the JMA record scaled at 0.8g.

63

Detrimental effects of urban tunnels on design seismic ground motions


G. Kouretzis1, G. Bouckovalas1, A. Sofianos1, P. Yiouta-Mitra1
1

National Technical University, Athens, Greece.

Abstract
This research deals with the issue of whether, and under what circumstances, the presence
of underground tunnels should be taken into account for the earthquake resistant design of
neighboring surface structures in urban areas. In order to investigate the effect of the
underground structure on surface seismic motion, a series of dynamic plane-strain
numerical analyses were conducted, considering a circular tunnel embedded in a
viscoelastic half-space, and a harmonic SV-wave excitation. The numerical methodology,
based on the Finite Difference Method, aims at quantifying the effect of the soil medium
characteristics, excitation frequency, tunnel diameter, depth of construction, and relative
flexibility of the lining compared to that of the surrounding soil. Conclusions include
preliminary criteria identifying the cases when the presence of an underground tunnel
should be considered in the design of a surface structure.

amplification of the free-field motion. The area


where this phenomenon is noticed is referred in
the literature as shadow zone.
- The presence of an underground tunnel
generates a parasitic vertical component of
surface motion, with significant amplitude.
The most important parameters, rising from
the analytical results presented by the aforementioned researchers to be influencing the
response at the ground surface, are:
- The ratio of the depth of structure axis
over the radius of structure, H/. For small
values of H/, not only the amplification due to
the underground cavity is larger, but also
seismic waves are trapped between the cavity
and the ground surface, thus producing more
complex response patterns.
- The dimensionless frequency n=2/,
where is the wavelength. Higher values of n
generally produce higher amplifications and
more complex response patterns.
- The dimensionless distance along the
ground surface from the projection of the
structure axis, x/, where is the radius of the
structure. The wave field affected from the

INTRODUCTION

The effect of the construction of underground structures such as Metro tunnels, on


aboveground buildings has been examined
thoroughly in the past, focusing mainly on the
resulting surface permanent settlements. Yet,
the effect that the presence of a massive
circular tunnel has on the seismic design of
nearby buildings is neglected in current
practice.
Nevertheless, many researchers (e.g. Lee
and Trifunac, 1979, Manoogian & Lee, 1996,
Manoogian, 1998, Lee & Karl, 1992, Barros &
Luco, 1993 and 1994, Lee et. al, 2001) have
investigated via analytical methods the effect
that the presence of a circular underground
structure, embedded in an elastic half-space,
has
on the
surface
ground
motion.
Summarizing their findings the following can be
said:
- Maximum amplification of the ground
surface motion, consisting of SV-waves, due to
the presence of an underground unlined circular
structure varies between 75% - 100%.
- The presence of a cavity can cause, under
certain conditions, intense and selective de-

64

Fig 1: Comparison of numerically computed normalized horizontal displacements on the ground surface, with
the results of Luco & De Barros (1994). The case shown here applies to =0.001, n=0.5, =1/3.
Table 1: Range of physical parameters corresponding to the numerical models.

Circular unlined tunnel

Dimensions
Shear wave velocity, Vs
Harmonic excitation
frequency, T

D = 520m

Circular lined tunnel

D = 520m, d = 0.040.17m
2001000 m/sec
0.05 1.0 sec

Code FLAC (Itasca, 1999). Following the


verification of the numerical model, a series of
parametric analyses are presented, aiming at
the definition of preliminary criteria identifying
the cases when the presence of an
underground tunnel should be considered in the
design of surface structures.

presence of the underground structure


attenuates at low rates, and is approximately
analogous to the inverse square root of the
distance from the tunnel axis. However,
analytical methods proposed in the literature
have produced results limited to x/<4.
- The relative stiffness of the tunnel lining,
compared to the surrounding soil, as stiffer
structures
appear
to
produce
lower
amplifications.
Existing analytical solutions provide a clear
overview of the problem, and identify the basic
parameters influencing the ground response.
Nevertheless, as they are based on complex
mathematics, they cannot be readily used for
parametric analyses to examine the effect of a
wide range of values of the aforementioned
parameters. For this reason, in the present work
the problem is treated numerically, with the aid
of the commercially available Finite Difference

NUMERICAL MODEL AND VERIFICATION

The numerical model is verified against the


analytical results of Luco and De Barros (1994),
who applied an indirect boundary integral
method based on two-dimensional Greens
functions to obtain the harmonic motion on the
ground surface. Their model consisted of an
infinitely long unlined cavity of circular crosssection with radius , embedded at a depth H,
oriented parallel to the free surface of the halfspace. The soil was modeled as a viscoelastic

65

The range for these parameters was based


on common characteristics of underground
tunnels and wave characteristics representing
high, as well as low frequency excitations.
Table 1 presents the range of the structure, soil
and motion characteristics examined in the
analyses. The combinations of these parameters results in a range of 0.051 for the
dimensionless
frequency
n,
while
the
dimensionless depth H/ obtains values of 17.
Obviously, since the depth of one radius would
apply to a tunnel emerging at the surface, it can
no longer be considered as underground. Thus,
it has been included only for reasons of
completeness.
The index of relative flexibility, J, was
calculated with the following equation (St. John
and Zahrah, 1987), considering a concrete
structure lining with Elining = 30 GPa and
lining = 0.2. Combined to the soil characteristics,
it yields two singular values, namely J=5 and
J=150, corresponding to a very stiff and a very
flexible tunnel, respectively.

medium, characterized by propagation velocities of P- and SV- waves equal to a and b


respectively, while the simplifying assumption
for hysteretic damping a=b= was also
adopted.
In this paper, the numerical model was
prepared along the above lines in order to
provide the means of verification. The
commercially available Finite Difference Code
FLAC (Itasca, 1999) was selected to perform
the analyses. The boundary conditions applied
at the artificial boundaries were a) lateral
dashpots to minimize wave reflections and
achieve free-field conditions, b) absorbing
boundaries at the bottom i.e. normal and shear
dashpots of coefficient c=CsVs (Kramer, 1996)
to represent the effect of radiation damping and
c) a stress boundary of amplitude xy=2CsVs at
the bottom to simulate the incoming harmonic
SV- wave. The factor of 2 in the above relation
corresponds to the fact that half of the input
energy is absorbed by the viscous boundary,
while in the above relations Cs is the velocity of
S-wave propagating through the soil medium,
is the soil mass density and Vs the excitation
particle velocity. Despite the existence of a
vertical plane of symmetry, the full model was
used due to the limitations of the numerical
code with respect to lateral dashpots. Hysteretic
damping is simulated in an approximate fashion
as Rayleigh damping according to the
guidelines of Itasca (1999) and Sofianos
(2003).
Figure 1 depicts a sample of the satisfactory
comparison with the Luco & De Barros
analytical model. In Figure 1, surface displacement amplitudes are normalized against the
amplitude of the free-field incident motion Us,
and they are illustrated against the normalized,
against the tunnel radius, distance from the
tunnel axis x/

J=

2
2Esoil (1 lining
)Rlining 3

Elining (1 + soil )t lining 3

(1)

Strong ground motion effects were


investigated for a distance of 18 radii from the
structure axis. Finally, a parameter that was
separately examined was the hysteretic
damping of the soil. Three values compatible
with the elastic model were selected, namely
=1%, 2% and 5%, corresponding to soft rockstiff soil formations. Results have shown no
practical difference between the three values,
therefore =2% was adopted in all analyses.
INDICATIVE RESULTS

A selection of diagrams is presented,


illustrating the normalized amplitudes of ground
surface displacements against the horizontal
distance from the structure axis of symmetry.
Normalization has been achieved by dividing
the results with the free-field horizontal
displacement amplitude, for horizontal but also
for vertical parasitic displacements, since there
is no free-field vertical component of motion to
compare with the latter.

PARAMETRIC ANALYSES

A series of parametric analyses were


conducted to quantify the effects of parameters
H/ (dimensionless depth), x/ (dimensionless
distance), n (dimensionless frequency) and J
(index of relative flexibility of the lining
compared to that of the surrounding soil), as
discussed in the introduction.

66

Fig. 2: Normalized horizontal displacement amplitudes for n=0.2, H/=2.

Fig. 3: Normalized horizontal displacement amplitudes for n=1.0, H/=2.

Fig. 4: Normalized vertical displacement amplitudes for n=0.4, H/=2.

further reduced to 15% (Fig. 6). It is therefore


confirmed that only for H/2 (Fig. 5) does any
strong motion aggravation arise, with maximum
amplification occurring at H/=1 and reaching
85%. Observations are somewhat different in
the case of the parasitic vertical component of
motion. At H/=1.5, maximum amplifications for
all frequencies are between 100-175 %.
Moreover, at a depth of three radii there is still a
70% amplification for n=0.6 (Fig. 7) and even at
four radii there is a 50% amplification for n=0.4.
The effect attenuates faster and drops at 15%
at a depth of seven radii.

Effect of the dimensionless frequency n

Low frequency values such as n=0.05 and 0.1,


that infer high values of wavelength as
compared to the tunnel dimensions, result in
waveforms practically unaffected by the
structure. For n higher than 0.2, the response
changes considerably. De-amplifications in the
shadow zone reach 55% and amplifications
30% (Fig. 2). For n=0.41.0 the shadow zone
becomes prominent (Fig. 3). Similar remarks
can be drawn for the vertical, parasitic,
component of motion. Higher frequencies result
in a vertical component that cannot be ignored
as in lower frequencies. Figure 4 depicts the
highest of these values at 90% of the free-field
horizontal component.

Effect of the relative flexibility of the tunnel J

For small depths, it can be observed from


Figures 8, 9 and 12 that decreasing values of J
tend to eradicate the effect of the presence of
the tunnel. At larger depths, the effect of the
tunnel lining is practically nullified.

Effect of the dimensionless depth H/

For depths larger than two radii, the


horizontal component amplification is no larger
than 25%. For depths larger than 7 radii it is

67

Fig. 5: Normalized horizontal displacement amplitudes for n=0.2, H/=1.5.

Fig. 6: Normalized horizontal displacement amplitudes for n=0.8, H/=7.

Fig. 7: Normalized horizontal and vertical displacement amplitudes for n=0.6, H/=3.

design of a nearby surface structure, all results


were evaluated via combined plots of the
maximum amplification values and of the
location of that maxima on the ground surface.
Figures 13 and 14 are only samples of such
plots, as the complete results are not shown for
brevitys sake. From Figure 13 it is drawn that
maximum amplification of the horizontal motion
reaches about 1.5 for a shallow (H/<2) unlined
tunnel. In Figure 14 it may be seen that for
small depths (H/<2), only the area close to the
tunnel is affected (0<x/<1.0). For 2<H/<4 the
affected area extends to 5 radii and for larger
depths it starts at 2 and extends to 7 radii.
Summarizing remarks from all relevant
diagrams, the following can be said for an
unlined tunnel:
- Ground response for excitations with
wavelengths larger than the tunnel diameter is
not affected by the presence of the tunnel.
- The presence of a tunnel results in

The last remark can be only cautiously


extended to all frequencies for the case of the
horizontal component of motion. As can be
seen from Figures 10 and 11, despite the large
depth of six radii, increasing frequencies result
in horizontal amplifications larger even than
those of the unlined tunnel. More specifically,
for J=5, when n=1.0 and H/=6.0, ANx fluctuates
from 0.45 to 1.38, while the same analyses for
an unlined tunnel resulted in amplifications of
0.54 to 1.17. With respect to the vertical
component however, the increase of the
structure stiffness always results in a decrease
of strong motion amplification, with the effect
becoming more prominent at smaller depths.
CONCLUSIONS

In order to extract preliminary criteria


identifying the cases when the presence of an
underground tunnel may be neglected in the

68

Fig. 8: Normalized horizontal displacement amplitudes for n=0.2, H/=2 for an unlined cavity (J=inf), very
flexible lining (J=150) and stiff lining (J=5).

Fig. 9: Normalized horizontal displacement amplitudes for n=0.2, H/=4 for an unlined cavity (J=inf), very
flexible lining (J=150) and stiff lining (J=5).

Fig. 10: Normalized horizontal displacement amplitudes for H/=6 n=0.4 for an unlined cavity (J=inf), very
flexible lining (J=150) and stiff lining (J=5).

Fig. 11: Normalized horizontal displacement amplitudes for H/=6 n=1.0 for an unlined cavity (J=inf), very
flexible lining (J=150) and stiff lining (J=5).

69

Fig. 12: Normalized vertical displacement amplitudes for n=0.4, H/=2 and 5 for an unlined cavity (J=inf), very
flexible lining (J=150) and stiff lining (J=5).

Fig. 13: Combined plot of maximum normalized horizontal displacement amplitudes due to the presence of an
unlined tunnel, for all depths and frequencies.

Fig. 14: Combined plot of the maximum amplification location, due to the presence of a flexible tunnel
(J=150), for all depths and frequencies.

70

Itasca Consulting Group Inc, (1999) FLAC, Fast


Lagrangian Analysis of Continua, Users
Manual, Mineapolis, MI, U.S.A.

amplifications that should be considered in the


design of a surface structure for H/ <3.0.
- The horizontal component of motion can
be amplified by 20% to 85% within a distance of
eleven radii from the tunnel axis (0<x/<11).
- The ground response is further
complicated by the appearance of a parasitic
vertical component of motion. Maximum vertical
displacement amplitudes occur within four radii
from the tunnel axis (0<x/<4), reaching values
of 1/3 to up to 3 times the value of horizontal
free-field component.
From the analyses for a lined tunnel, the
following remarks can be drawn:
- Stiffer linings generally attenuate the
effect of the tunnel presence on the ground
response.
- For n>0.8 and H/>6.0, stiffer tunnels
cause unexpected for such depths amplifications of about 30%.
- The horizontal component of motion can
be amplified from 20% up to 40%.
The vertical component of motion reaches
values from 1/3 to 1.5 times the value of the
horizontal free-field component.
- The value of the relative stiffness does
not alter the shadow zone, which therefore
coincides with that of the unlined tunnel.

Lee, V. W., (1977) "On the deformations near


circular underground cavity subjected to
incident plane SH-waves ", Proc. Application of
computer methods in engineering conference,
University of South California, L.A., pp.951-962.
Lee, V. W., Karl, J., (1993) "On the deformations
near a circular underground cavity subjected to
incident plane P waves", European Earthquake
Engineering 1, 29-39.
Lee, V.W., Karl,J., (1992) "Diffraction of SV-waves
by underground, circular, cylindrical cavities.",
Soil Dynamics and Earthquake Eng. 11,445456.
Lee, V.W., Trifunac, M.D., (1979) "Response of
tunnels to incident SH- waves.", J.eng.
mech.div. ASCE 105, 643-659.
Luco, J.E., De Barros F.C.P., (1994) "Dynamic
Displacements and stresses in the vicinity of a
cylindrical cavity embedded in a half- space.",
Earthquake Eng.& Structural Dynamics
23,321-340.
Manoogian ,M. E., "Surface motion above an
arbitrarily shaped tunnel due to elastic SH
waves." in Proc. Geotechnical Earthquake
Engineering and Soil Dynamics III.
Manoogian, M.E., Lee, V.W., (1996) "Diffraction of
SH-waves by subsurface inclusions of arbitrary
shape", J.eng. mech.div. ASCE 122,123-129.

ACKNOWLEDGEMENTS

The research was funded by Earthquake


Planning and Protection Organization, Greece.

Penzien.J., (2000) "Seismically induced racking of


tunnels linings", Earthquake Eng.& Structural
Dynamics 29,683-691.

REFERENCES
Kramer L. S., (1996) Geotechnical Earthquake
Engineering, Prentice Hall.

Sofianos, A.I., (2003) Effect of underground


structures on design response spectra, Final
Report (3), Earthquake Planning and
Protection Organization, Greece.

Datta, T.K., (1999) Seismic response of buried


pipelines: a state-of-the-art review, Nuclear
Engineering and Design, Vol. 192, Issues 2-3,
271-284.

St. John, M.C., Zahrah, T.F., (1987) "Aseismic


Design of Underground Structures", Tunneling
& Underground Space Technology 2(2),165197.

Davis, C.A. , Lee, V. W. , Bardet, J.P., (2001)


"Transverse response of underground cavities
and pipes to incident SV waves.", Earthquake
Eng.& Structural Dynamics 30, 383-410.

Vanzi,V., (2000) "Elastic and inelastic response of


tunnels under longitudinal earthquake excitation.",
J. Earthquake Engineering 4 (2),

De Barros F.C.P. & Luco, J.,E., (1993) "Diffraction of


obliquely incident waves by a cylindrical cavity
embedded in a layered viscoelastic halfspace", Soil Dynamics and Earthquake Eng.
12, 159-171.

71

Seismic Analysis of Near-Surface Tunnel Cross-Sections


Christos Vrettos
Technical University of Kaiserslautern, Germany

Abstract
Seismic loads exerted on tunnels in soft soils are controlled by the vibrational characteristics
of the surrounding soil and the flexibility of the tunnel lining. The analysis is performed by
the deformation method using imposed displacements. Available closed-form expressions
to calculate deformations and sectional forces are compared. By means of a finite element
model the accuracy of the decoupled analysis of the associated soil-structure-interaction
problem is assessed, and the influence of the tunnel embedment depth is investigated.
Finally, a 2D dynamic analysis of a shallow and of a deep tunnel in a layered soil deposit is
carried out to identify the relevant features of the seismic response.

INTRODUCTION

the tunnel lining. The present paper deals with


this loading case.

Tunnel structures are generally considered


less vulnerable to seismic actions than above
ground structures. In contrast to surface
structures, whose seismic response is governed
by inertia effects, the response of tunnels
embedded in the ground is primarily kinematic,
i.e. it is caused by the compatibility of the tunnel
deformation to that of the surrounding ground.
Therefore, soil-structure interaction effects are
of fundamental importance. The stiffness ratio
between tunnel structure and soil determines
the intensity of sectional forces. The analysis is
usually carried out by imposing appropriate
displacement patterns (deformation method).
Inertia effects are of secondary importance and
can be neglected. The available analysis
methods are presented by St. John and Zahrah
(1987), Wang (1993), Hashash et al. (2001),
and also by the author (Vrettos, 2005).
Impinging seismic waves cause in the
longitudinal direction i) axial deformations due
to alternating compression and tension, and ii)
bending due to curvature arising from the wave
components with particle motions perpendicular
to the tunnel axis. These cases are not
considered herein.
In the tunnel cross section vertically
propagating shear waves cause distortion of the
circular or rectangular cross sectional shape of

OVALING OF CIRCULAR TUNNELS

We first consider circular tunnel cross


sections subjected to shearing under plane
strain conditions. The equations given in the
following are based on the work by Wang
(1993) and Penzien (2000).
When the tunnel stiffness is equal to that of
the ground, the tunnel is modelled as a hole
filled with soil. The diametric strain ' ff in the
tunnel for shear strain J is
' d ff
d

J
2

(1)

where d ist he tunnel diameter.


For very low values of the tunnel stiffness
we obtain the other limiting case of a perforated
soil slice. The diametric strain 'dc is then

'd c
d

r 2 J (1  Q)

(2)

where Q is the Poissons ratio of the linearelastic soil medium. It becomes evident that the
relative stiffness between tunnel and soil
dominates the response. Two quantities are

72

used for this purpose: the compressibility and


the flexibility ratio, C and F, respectively:
C

E (1  Q "2 ) d
2 E " t " (1  Q)(1  2Q)
Q 2" ) d 3

E (1 
48 E "I " (1  Q)

K2

F [(1  2Q)  (1  2Q)C ]  (1  2Q)2 / 2  2


F [(3  2Q)  (1  2Q)C ]  C[2.5  8Q  6Q 2 ]  6  8Q
(10)

(3)

Expressions for diametric strain and bending


moment have not been derived yet, and it is
recommended to apply for the design the
values for full-slip conditions.
Solutions
for
sectional
forces
and
deformations due to soil-structure interaction
have also been derived by Penzien &
Wu (1998) and Penzien (2000). Assuming fullslip contact conditions the following closed form
solutions are obtained:

(4)

In the above equations E is the Youngs


modulus of the soil, and E" , I " , Q " and t " are the
Youngs modulus, the moment of inertia, the
Poissons ratio, and the thickness of the tunnel
lining.
Analytical solutions for the diametric strain,
the axial load, and the bending moment have
been derived by assuming full-slip contact
conditions between tunnel lining and soil.
Following Wang (1993) we obtain:

' d tunnel
' d ff

Nmax

2
K1F
3

n
' dtunnel

(5)

E d
1
J
K1
6
(1  Q) 2

1

(6)

Nmax

Mmax

Qmax

Mmax

E d
1
K1
J
6
(1  Q) 4

Rn
12(1  Q)
2F  5  6Q

Nmax

6E " I " R n

d 2 (1  Q 2" )
3E "I " R n

d (1  Q 2" )

d
J
2

(12)

12E "I " R n

d 2 (1  Q 2" )

(11)

(13)
J

(14)

4(1  Q)

(15)

Dn  1

(8)
with

Sliding along the contact surface is possible


only in tunnels embedded in very soft soils
under strong seismic actions. In most cases in
practice an intermediate condition between fullslip and bonded contact will develop. The
assumption of sliding yields conservative
estimates of sectional strain and bending
moments.
Under seismic shear loading maximum
values for axial forces in the tunnel are obtained
under the assumption of bonded contact:
E d
rK 2
J
(1  Q) 2

Rn

with R n defined as the racking ratio for fullslip conditions:

(7)

with
K1

R n ' d ff

Dn

12E "I " (5  6Q)

(16)

d 3G(1  Q 2" )

where G denotes the shear modulus of the


soil.
For the case of bonded contact the
corresponding solutions are obtained by
replacing R n by R in the above equations (11)
to (14), cf. Penzien (2000):
' d tunnel

(9)

Nmax

where

73

R ' d ff

12E "I " R

d 2 (1  Q 2" )

d
J
2

(17)
(18)

M max

Qmax

3E "I " R
Q 2" )

d (1 
12E "I " R
2

d (1 

Q 2" )

NUMERICAL INVESTIGATION ON
OVALING OF CIRCULAR TUNNELS

(19)
J

The numerical study is carried out by using


the finite element method code PLAXIS (2004).
Since modelling of full-slip is difficult with this
code a bonded contact is assumed here. The
equations presented above for a perforated soil
slice with/without tunnel are valid for a fullspace, i.e. the boundaries are located in a
sufficiently large distance from the structure.
The required minimum size of the FEM-mesh is
determined by comparing the numerical solution
with the analytical solution of a perforated soil
slice equation (2). For the soil assumed here
( Q 0.3) and J = 2.5210-3, we obtain for a
quadratic mesh of dimension equal to 15d an
accuracy in the value of diametric strain of
approx. 1.3%. The respective results for the
racking ratio and the sectional forces are:

(20)

where
R

4(1  Q)
D 1
24E "I " (3  4Q)
r

d 3G(1  Q 2" )

(21)
(22)

Comparison of the above equations is made


by using an example. The seismic shear strain
is estimated from the approximation by
Newmark (1967):
J

v
Cs

(23)

R 2.164,
Mmax 157.3 kNm/m,

where v is the maximum particle velocity of


the seismic excitation, Cs
G / U is the shear
wave velocity in the soil, and U is the density.
The following numerical values are selected
here: v = 0.63 m/s, Cs = 250 m/s, and
U = 1.92 Mg/m3. For the Poissons ratio we set
Q = 0.3.
For the tunnel lining the following values are
selected: d = 6 m, t " 0.3 m (cross sectional
0,3m2/m and moment of inertia
area A"
E"
I"
0.00225 m4/m),
24.80106 kPa,
Q " 0.20. It is further assumed that the tunnel
is located in a large depth so that the influence
of the free ground surface can be ignored.
For the case of full-slip equations (6) - (7)
and equations (12) - (13) yield identical results:
Nmax 62.94 kN/m und Mmax 188.81 kNm/m.
For the case of bonded contact equation (9)
by Wang (1993) yields an axial force
Nmax 1.045.8 kN/m while equation (18) by
Penzien (2000) yields a much lower value of
Nmax 124.4 kN/m. Bending moment and shear
force from equations (19) and (20) are
Mmax 186.95 kNm/m and Qmax 124.64 kN/m,
respectively. Racking ratio R according to Wang
is 2.58, while the equation by Penzien yields
2.55. The large discrepancy between equations
(9) and (18) has also been identified by
Hashash et al. (2005) and is investigated next.

Nmax
Qmax

1.040 kN/m,
108.1 kN/m

Based on this, the solution by Penzien


underestimates the axial force N by a factor of 8
and is thus not acceptable for use in design
practice. In contrast the solution by Wang
shows an excellent agreement with the
numerical solution. The values for the other
sectional forces show an accuracy of approx.
20%. Similar findings have been reported by
Hashash et al. (2005).
RACKING OF RECTANGULAR TUNNELS

Closed-form solutions for sectional forces


considering also the effects of soil-structure
interaction do not exist. A pseudo-coupled
procedure is usually applied: The distortion of
the structure is first estimated and subsequently
imposed on the structural system in order to
determine the sectional forces, cf. Wang (1993)
and Penzien (2000). This procedure is
summarized in the following.
The maximum relative deformation of the
tunnel lining due to the shear strain J induced
by the earthquake is:
' tunnel

J tunnel H

(24)

where H is the height of the rectangular


tunnel box. The relative free-field displacement

74

when the presence of the tunnel is ignored is


' ff J H .
Depending on the value of the relative
stiffness between tunnel and surrounding soil
' tunnel can be larger or smaller than ' ff . The
corresponding racking ratio is defined by:

' tunnel
' ff

Multiplication by R yields 'tunnel . This


displacement is finally applied to the structural
system of the tunnel, and the sectional forces
are determined.

(25)

The relative stiffness between tunnel and


soil is defined in terms of the flexibility ratio F:

GB
k" H

(26)
Fig 1:

where G is the strain-compatible shear


modulus of the soil, B is the width of the tunnel,
and k " is the horizontal stiffness of the
rectangular box. The latter is determined by
standard structural analysis methods on a
simple system consisting of the tunnel box
without the surrounding soil. It should be
noticed that the definition of the flexibility ratio is
different for circular and rectangular tunnels, cf.
equation (4) and equation (26), respectively.
A closed form solution for the racking ratio R
for deeply embedded tunnels has been derived
by Penzien (2000). It is determined from
equation (21), where D is given by the
following relationship:
D

ks

k " (3  4Q)
ks

G
H

Simple system to determine horizontal


stiffness of tunnel lining; W " k " / B .

The accuracy of the above procedure is


examined by comparison with a numerical
solution. A point that deserves attention is the
proper selection of Youngs modulus for the
tunnel under plane strain conditions. Finite
element codes like PLAXIS directly use the
Youngs modulus E and the Poissons ratio Q ,
while in beam analysis E has to be replaced by
E E /(1  Q 2" ) .
As an example we consider a rectangular
tunnel with dimensions B / H = 10 / 4 m, wall
thickness t" = 0.30 m, E" = 24.80106 kPa, and
Q " 0.20. For the soil we select G = 120 MPa
and Q = 0.3. The seismic shear strain is
assumed to act statically with J = 2.5210-3.
Considering the above comments regarding
the Youngs modulus E, the flexibility ratio F for
equation (26) can be calculated via an
expression given by Wang (1993) to F = 48.17.
PLAXIS yields for the system depicted in Fig 1
under unit shear stress loading and E A f a
maximum horizontal displacement equal to
1.6110-3 m, that corresponds to F = 48.30.
The racking ratio R according to
Penzien (2000) is then calculated with
D = 0.0373 to R = 2.70. For determining the
sectional forces in the lining the tunnel box
structure is subjected to a differential
displacement between roof and base equal to
' tunnel = 2.702.5210-34 = 27.2210-3 m.
In order to check the accuracy of this
pseudo-coupled procedure we compute the

(27)
(28)

The racking ratio is equal to 1 when the


stiffness of the tunnel lining k " is equal to that
of the soil k s . When k " becomes very small,
we obtain the limiting case of a perforated fullspace slice, i.e. R 4(1  Q) . It is evident, that
for very flexible structures the distortion of the
structure may be larger than that in the freefield, i.e. R ! 1.
Free-field displacement ' ff is determined
from a seismic site response analysis of the
particular soil profile or estimated by the
Newmark
approximation,
equation (23).

75

2D SEISMIC RESPONSE ANALYSIS

response of the entire tunnel-soil system under


shear loading using the code PLAXIS. The size
of the mesh is set equal to 10 times the tunnel
width and height, respectively. Distortion of the
tunnel is calculated from the lengths of the two
diagonals d1 and d 2 of the distorted cross
section:
' tunnel
J tunnel
(d 22  d12 ) /(4BH )
(29)
H

A dynamic 2D analysis by means of the


finite element method as usually carried out in
the frame of the detailed design of tunnel
projects is presented next.
We consider a layered soil deposit with a
thickness of 40 m subjected to a seismic motion
defined at outcropping rock. The acceleration
time history selected refers to the Loma Prieta
1989 earthquake recorded at Palo Alto SLAC
Lab, USGS Station 1601. The site conditions
correspond to rock. In the predominant
horizontal direction peak ground acceleration is
0.278 g that is scaled here to a peak value of
0.27 g.
First, a 1D free-field seismic site response
analysis is conducted using SHAKE. For the
variation of shear modulus and damping ratio
with shear strain we assume for all layers a
clayey soil of medium plasticity. The initial
small-strain shear moduli as well as the straincompatible ones are depicted in Fig 2. At the
interface between soil and bedrock a maximum
acceleration of 0.20 g is obtained from the
analysis while at the soil surface the maximum
acceleration is 0.45 g. The distribution of
maximum shear strain with depth is given in the
top graph of Fig 3.

From the numerical analysis we get J tunnel =


6.510-3 yielding a racking ratio R = 2.579.
Hence, the accuracy compared to the closedform solution by Penzien (2000) is 5 %.
INFLUENCE OF EMBEDMENT DEPTH IN
APPROXIMATE SOLUTION

The above closed form solutions are valid


for deep tunnels. The presence of a free
surface results in shear strain that varies with
depth. Furthermore, the stress distribution in
the vicinity of the tunnel depends on the
distance between tunnel roof and ground
surface. The effect of tunnel embedment depth
is investigated next by means of a numerical
analysis. We consider the rectangular tunnel
cross section used in the previous example and
vary the position of the tunnel relative to the
ground surface with cover thickness ranging
from 18 m to 2 m. The distance of the tunnel
from the bottom mesh boundary is kept
constant. The applied shear strain is set equal
to J = 2.52 10-3 m. The computed values of the
racking ratio R according to equation (25) are
summarized in Table 1.

Depth [m]

-10

Table 1: Variation of racking ratio with soil


cover thickness
tcover
[m]

18

12

2.58

2.52

2.44

2.26

2.08

-20

-30

It can be seen that the racking ratio varies


only weak with the soil cover thickness t cov er .
For a cover thickness equal to the tunnel height
the value of racking ratio is reduced by 13%
compared to the value of a deep tunnel. Similar
results are obtained for a much softer soil.
Hence, the solutions for deep sited tunnel
consists a practicable approximation.

-40
0

100

200

300

Shear modulus [MPa]

Fig 2:

76

Shear modulus profile in 1D analysis: The


dotted line is for the initial values, the solid
line for the strain compatible ones.

For the tunnel we select the single cell


structure of the previous examples with
dimensions height / width equal to 4 m / 10 m.
Two
positions
will
be
investigated
corresponding to soil cover thickness of 16 m
and 5 m, respectively.
The 2D seismic response analysis is carried
out using PLAXIS. The system width is set
equal to 100 m. The code works in the timedomain using frequency dependent Rayleigh
damping for the soil, thus requiring some
approximation to simulate the strain-compatible
hysteretic damping obtained from the 1D
analysis. For the shear modulus we use
constant values for the individual layers as
given in Fig 2. For the Poissons ratio we
assume Q = 0.3 for all layers. The seismic input
is defined at the bottom boundary of the finite
element mesh and corresponds to the time
history obtained by SHAKE at the soil/bedrock
interface with a maximum acceleration of
0.20 g.
The free-field analysis of this 2D system
yields at the surface a maximum acceleration of
0.3 g that is considerably lower than the 1D
value. The source of this discrepancy may be
found mainly in the 2D nature of the model, but
also in the Rayleigh damping approximation.
In the next step we analyze the system
including the tunnel structure. We monitor the
response at the upper right corner of the tunnel
structure, point C. The sectional forces are
evaluated at that point in time when the
maximum acceleration at point C occurs. For
the deep tunnel the maximum bending moment
is 565.2 kNm/m at an acceleration in point C
equal to 1.57 m/s2. For the shallower tunnel we
obtain a much lower value for the bending
moment equal to 283.3 kNm/m while the
acceleration at point C shows a much higher
value of 2.8 m/s2, cf. Fig 4.
Under 2D free-field conditions the
acceleration values at the depth level
corresponding to point C are: 1.72 m/s2 for the
deep and 2.45 m/s2 for the shallow tunnel,
respectively.
The fact that although the acceleration for
the deep tunnel is lower the associated bending
moment is larger is attributed to the larger shear
strain level along the height of the tunnel, cf.
Fig 3. This indicates that relative deformation is
governing the tunnel response.

Depth [m]

-10

-20

-30

-40
0

0.05

0.10

0.15

0.20

max. shear strain [%]

Depth [m]

-10

-20

-30

-40
0

0.05

0.10

0.15

0.20

shear strain [%]

Fig 3: Distribution of maximum shear strain with


depth. The top graph is from the 1D, the
bottom one from the 2D free field analysis
at the time of maximum acceleration at the
level of the shallow tunnel. The dotted lined
boxes indicate the tunnel position of the two
cases considered.

77

Hashash, Y.M.A., Hook, J.J., Schmidt, B., and Yao,


J.I.C. (2001) Seismic analysis and design of
underground
structures,
Tunnelling
and
Underground Space Technology, Vol. 16, pp.
247-29.
Newmark, N.M. (1967) Problems in wave
propagation in soil and rock, Proc. Int. Symp.
Wave Propgation and Dynamic Properties of
Earth Materials, New Mexico, Univ. of New
Mexico Press.
Penzien, J. (2000) Seismically induced racking of
tunnel linings, Earthquake Engng. Struct. Dyn.,
Vol. 29, pp. 683-691.
Penzien, J., and Wu, C.L. (1998) Stresses in linings
of bored tunnels, Earthquake Engng. Struct.
Dyn., Vol. 27, pp. 283300.
PLAXIS b.v. (2004): PLAXIS Finite Element Code for
Soil and Rock Analyses, Version 8, Delft,
Netherlands.
St. John, C. M. and Zahrah, T. F. (1987) Aseismic
Design of Underground Structures, Tunnelling
and Underground Space Technology, Vol. 2,
No. 2, pp. 165-197.
Fig 4:

Distribution of bending moments at the


point in time when maximum acceleration
occurs at the upper right connection
roof/wall.

Vrettos, C. (2005) Earthquake resistant design of


tunnels,
Rational
Tunnelling,
2nd
Summerschool, Innsbruck, D. Kolymbas &
A. Laudahn (Eds), pp. 261-283.
Wang, J.-N. (1993) Seismic Design of Tunnels: A
State-of-the-Art Approach, Monograph 7.,
Parsons, Brinckerhoff, Quade and Douglas Inc.,
New York.

CONCLUSIONS

Available closed form expressions for


computing sectional forces and deformations
due to ovaling/racking deformation of tunnels
have been compared with solutions obtained by
a finite element code, and some known
discrepancies for the condition of bonded
contact have been validated. The effects of the
free soil surface are assessed.
The difficulty in matching 1D and 2D free
field analyses is shown. The seismic response
of a single cell tunnel embedded in layered soil
is investigated for two embedment depths
demonstrating the relevance of kinematic soil
structure interaction and the appropriateness of
the deformation method for the design.
REFERENCES
Hashash, Y.M.A., Park, D, and Yao, J.I-C. (2005)
Ovaling deformations of circular tunnels under
seismic loading, an update on seismic design
and analysis of underground structures,
Tunneling and Underground Space Technology,
Vol. 20, pp. 435-441

78

Sliding of Rigid Block on Sloping Plane:


The Surprising Role of the Sequence of Long-Duration Pulses
E. Garini and G. Gazetas
National Technical University of Athens, Greece

Abstract
A numerical study is presented for a rigid block supported through a Coulomb friction
contact surface on an inclined plane subjected to parallel excitation. As excitation we utilize
idealized wavelets and near-fault seismic records strongly influenced by forward directivity
or fling-step effects (Northridge 1994, Kobe 1995, Kocaeli 1999, Chi-Chi 1999). For a given
level of peak acceleration the sliding displacement is found to be dependent on the following
parameters: the ratio ac1/aH of the critical sliding acceleration ac divided by the peak input
acceleration aH, the excitation frequency fH, the inclination angle of the plane, and, as
significantly, on the sign (+ or -), the sequence, and even the details of pulses of the
excitation time history. The parametric analysis shows that the displacements induced due
to sliding on inclined plane may exceed by far those induced on a horizontal plane, due to
the accumulation of slippage in one direction. Dynamic response of the block-base system
is found to be very sensitive to the unpredictable details of the excitation time history.

INTRODUCTION

acceleration aH (in g units). Moreover the rigid


block is supported through a Coulomb frictional
surface with a constant coefficient of friction
on the inclined plane.
The forces acting on the block are, as
illustrated in Figure 1b: (a) the weight, B, with its
two components, (b) the inertial force, F, due to
the blocks acceleration, (c) the perpendicular
reaction, Fk , of the plane and (d) the friction
force, T = Fk. Dynamic equilibrium of the block
results to the yielding downhill or uphill
acceleration. Respectivelly:

In earthquake geotechnical engineering the


analog of dynamic sliding of a block on an
inclined plane has been used to estimate the
response of earth dams, embankments, and
retaining walls during earthquakes. Introduced
in 1965, by Newmark the analog was further
utilized Seed and Martin (1966), Ambraseys
and Sarma (1967), Makdisi and Seed (1978)
who developed procedures for predicting the
permanent displacements of dams, applying
Newmarks sliding-block concept. Additional
studies of slip in relation to the strength of
ground shaking have been made by Richards
and Elms (1979) for evaluating the response of
shallow foundations and gravity retaining walls.
Also, Whitman and Lin (1983), Yegian et al
(1988), Gazetas and Uddin (1995), Kramer and
Smith (1997) used Newmarks model in
analyzing the stability of dams and
embankments.
The seismic slope stability problem is
analogous to the problem of a rigid block of
mass m, resting on an inclined plane which
forms an angle with the horizontal (Fig 1a).
The plane is subjected to seismic excitation
parallel to the base with peak ground

ac1 = cos sin


ac2 = cos + sin

(1)
(2)

Evidently, ac1 < ac2.


The acceleration ac1 at which the mass yields
downhill is called critical sliding acceleration.
Imposed bases acceleration in every moment
has a different value, aH(t), which can be
greater of lower than ac1. As long as the ground
acceleration, aH(t), does not exceed ac1 the
block does not slip. When aH(t) surpasses ac1
sliding takes place. The block can no longer
accelerate as fast as the plane and moves
relative to the base with a constant yielding
acceleration ac1. Because the block is assumed

79

cycle at the Ricker wavelet. As the ratio ac1/aH


decreases from 0.4 to 0.2, these two smaller
pulses cause significant sliding in the opposite
direction of the main cycle. For the other types
of idealized wavelets (T-Ricker, Sinus, Tsang)
such paradoxical behavior is not noticed (see
Figure 3).
Slippage is roughly inversely proportional to
ratio ac1/aH. By increasing ac1/aH, the velocity
gradient of the block is increasing and
consequently sliding occurs for a shorter period
with duration almost proportional to ac1/aH. The
explanation is conspicuous as can be seen
from Figure 4: In general, the resulting sliding is
equal to the area which is enclosed between
the velocity time histories of the block and the
base. The first sliding uphill is almost the same
for different values of ac1/aH (striped area in
velocity time-histories of Fig.4). However, at t =
1.7 sec the block stopped sliding uphill and
starts yielding at the opposite direction. That
makes the difference at the blocks response:
As illustrated in Fig. 4, when ac1/aH = 0.05
sliding downhill lasts 2 sec (starts at 1.7 sec
and finishes at 3.7 sec) whereas for ac1/aH = 0.1
yielding downhill lasts only 1 sec (from 1.7 sec
until 2.7 sec). As a result, the induced
displacement at the end of the sliding, which is
simply the shaded triangular area in Fig. 4, is
reduced to one half when the ratio ac1/aH is
doubled.

to be rigid, the acceleration throughout sliding is


constant. The movement continues until the
velocities of the block and the ground equalize.
Knowing the critical acceleration and the time
history
of
base
excitation,
permanent
displacements in every sliding period can be
calculated by a straightforward double
integration process.
Due to the transient nature of the
earthquake loading, the block may be subjected
to a number of acceleration pulses equal to, or
higher than its critical acceleration, which simply
produce some permanent deformation rather
than complete failure. Theoretically, sliding on
an inclined plane can happen in both directions:
uphill or downhill. However in practice, for
planes with inclination angle greater than 5o
sliding occurs only at downhill (where the mass
yields with acceleration ac1). Therefore, the
permanent deformation induced in every sliding
period, is added to the displacements of the
previous sliding periods.
ASYMMETRIC SLIDING INDUCED BY
IDEALIZED WAVELETS

Four types of idealized wavelets are used as


base excitation in this study: i) Ricker wavelets
of characteristic frequency: fH = 1, 2, and 4 Hz,
ii) One-cycle sinus pulse of frequency: fH = 1.5
and 3 Hz, iii) T-Ricker wavelet of frequency: fH
= 1 and 2 Hz, and iv) Tsang wavelet of
frequency: fH = 1 and 2 Hz. A number of
numerical analyses are performed in order to
understand the influence of each parameter in
asymmetric sliding. So, a wide range of values
is given to each parameter: ac1/aH = 0.05, 0.1 ,
0.2 , 0.4 , 0.6 and = 0o, 5o, 25o. Every
excitation time history is applied with its normal
and opposite sign (right and left in Figures).

Influence of the excitation frequency fH


Increasing the excitation frequency, leads to
dramatically decreasing slippage (with the ratio
ac1/aH and the angle remaining constant). The
explanation is straightforward: The higher the
frequency, the shorter the duration of block
acceleration relatively to the ground (Fig. 5).
Consequently, the velocities of the block and
the ground equalize more quickly and sliding
occurs for a shorter period, hence smaller
displacement. In Figure 6 is depicted the sliding
response for two different frequencies of Tsang
pulse in respect to yielding acceleration ratio
ac1/aH. The aforementioned trends are seemed
to be valid: the higher the frequency, the
smaller the slippage.
In addition, sliding is inversely proportional
to the square of the vibration frequency. Figure
6 shows that when frequency fH is doubled, the
induced slippage reduces by a factor of 4.

Influence of ac1/aH ratio


Evidently, increasing the ratio ac1/aH while
keeping the values of inclination angle and
peak ground acceleration aH constant, leads to
reduced slippage. Sliding displacements are
decreasing monotonically with increasing
friction coefficient, when > 0. However, when
= 0o (i.e. for sliding on horizontal plane) for a
Ricker excitation a paradoxical reduction of is
observed with decreasing ac1/aH (Fig. 2). This
phenomenon is caused by the presence of the
two smaller pulses before and after the major

80

The relative displacement d(t) results from the


integration of the differential of velocities
between the block and the base. In simple
cases such as in Fig. 6, this integral is given by
the area of a fundamental shape. Particularly in
Fig. 6, is the area or the almost rightangled
triangle whose one side is the yielding period
and the other one is the blocks velocity at the
beginning of the sliding. Notice that for fH = 1.5
Hz sliding lasts 4 sec (from 1.7 to 5.7 sec) while
for fH = 3 Hz sliding holds the half time (from 1.3
to 3.3 sec). Moreover, blocks velocity at the
beginning of the yielding is 2.2 m/s when fH =
1.5 Hz but it is reduced to the one half (1.1 m/s)
when fH = 3 Hz. Therefore by doubling the
excitation frequency, response increased by a
factor of 4.
Normalization of asymmetric response with
respect to the ratio of peak ground acceleration
divided by the square of peak ground velocity,
aH/VH2, converges the results to a single line; for
idealized pulses like T-Ricker and sinus, see
Figure 7 and 8 respectively. Observe that for all
excitation frequencies, fH, the normalized
slippage for any particular ratio ac1/aH is of the
same value.

Influence of excitation direction


Response of sliding block is strongly
depended on the direction (+ or -) of the
imposed excitation, as illustrated in Fig. 11, only
for the cases of asymmetric sliding ( 0).
A dramatic example is offered in Fig. 12. It is
presented the response of a rigid block sliding
on an inclined plane subjected to a single sinus
pulse. Time-histories on the left shows that the
first positive acceleration pulse of the base
induces downhill sliding, reaching a mere 0.6 m
in magnitude for 0.7 sec duration. The
immediately ensuing negative acceleration
pulse decelerates the body until sliding stops
without reversing uphill.
In contrast, when the direction of sinus pulse
reverted (right plots in Fig. 11) the sequence of
positive and negative cycles change. This time
the first acceleration pulse is the negative one,
causing as expected only a minor slippage
uphill, hardly moving the block from its rest
position. Then the positive acceleration pulse is
imposed leading to an almost unstoppable
downhill sliding of 4.2 m for 4 sec duration! It
appears that the vast difference between these
two cases is fully justified by the crucial
presence of the negative acceleration pulse at
the beginning or the end of excitation, because
this pulse acts like a brake in blocks sliding. A
practical conclusion: during earthquakes,
observed differences in behavior between
various
slopes,
retaining
walls,
and
asymmetrically-yielding structures (in general),
usually attributed to local site effects or different
structural characteristics, may only be the
unfortunate outcome of asymmetric sliding or
yielding,
arising
from
the
completely
unpredictable lack of symmetry of ground
shaking.

Influence of inclination angle


Generally, as the inclination of the base
increases, both the magnitude and the duration
of downward sliding also increase. As angle
increases, while keeping ac1/aH and fH constant,
sliding uphill becomes harder. It can be seen
from Figure 9 that for = 5o the block after the
first sliding downhill, starts at the moment t = 2
sec sliding uphill while for = 25o at the same
time (t = 2 sec) block is moving in absolute
contact with its baseno uphill sliding occurs.
Figure 10 depicts the acceleration, velocity
and slippage response of a block on either a 5o
inclined plane (left) or a 25o inclined plane
(right). The excitation is a Tsang pulse of 1 Hz
frequency. In case of the 5o slope not only
downward movement occurs; mild inclination
permits an uphill sliding of almost 0.3 m. On the
other hand, when the slope is steep ( = 25o)
yielding is always downhill leading to
substantially larger slippage.

ASYMMETRIC SLIDING INDUCED BY NEARFAULT GROUND MOTIONS

To verify the above trends obtained only


with idealized wave forms, analyses with actual
accelerograms are performed. Fortunately in
the last two decades, several strong
earthquakes have occurred providing a great
number of near-fault records, which are
influenced by forward rupture directivity and
fling-step effects (Somerville 2000, Abrahamson
2003). A selection of 15 horizontal components
of records is used as input shaking parallel to

81

the significance of the form of accelerogram is


portrayed in Figures 15, 16 and 17, which refer
to the north-south component of TCU068
record of the Chi-Chi 1999 earthquake. This
record comprised long period pulses of T 3
sec which reach acceleration values of only
0.20 g (compared with the much higher PGA
value of 0.42 g).
As can be seen from Figure 15, for a ratio
ac1/aH = 0.05, this excitation produces a sliding
displacement of 25 m! This is clearly due to the
mentioned long period pulse which starts at
about 7 sec. To further prove the importance of
this long duration pulse, a simplifying
approximation of the TCU068 n-s record is
made by clipping-off high accelerations and
keeping only the fundamental pulses of the
record. Figs 16-17 reveal that the resulting
slippage remains almost the same with that
produced with the actual record, especially in
case of the inverted accelerogram.

the base:

JMA 00, JMA 90 (Kobe 1995)


Fukiai (Kobe 1995)
Takatori 00, Takatori 90 (Kobe 1995)
Takarazuka 00, Takarazuka 90 (Cobe
1995)
No 4-140, 4-230, 6-140 (Imperial Valley
1979)
TCU068 NS, TCU068 EW, TCU102 NS
(Chi-Chi 1999)
Lucerne 275 (Landers 1992)
Sakarya EW (Kocaeli 1999)
Yarimca 60 (Kocaeli 1999)
Pacoima Dam 164 (Northridge 1994)
Santa Monica 90 (Northridge 1994)
UCLA 360 (Northridge 1994)
Rinaldi 228 (Northridge 1994)
Jensen 22 (Northridge 1994)
Newhall 360 (Northridge 1994)

For every record a series of numerical analyses


is performed and the influence of each
parameter in asymmetric sliding is assessed.
The following trends are noticed in the results of
all time-history analyses (not only those shown
here).

Influence of inclination angle


Figure 18 shows that when = 5o, the block
yields downhill as well as uphill (for ratio ac1/aH
= 0.05, one major uphill sliding occurred at t =
6.1 sec) while for larger values, block slides
exclusively downhill. Nonetheless, the induced
displacements are roughly the same. As a
consequence, few slidings happened uphill
even if the input motion includes multi-cycle
pulses such as in JMA, Fukiai, and Takatori
records.
Contrasting difference observed between
the symmetric ( = 0) and the asymmetric (
0) sliding response, as indicated by Figures 19
and 20. The key point is that in case of
horizontal plane, the block can slide both ways;
thus slippage induced in one direction during a
cycle can be inverted in next ones returning the
block to its initial position.

Influence of ac1/aH ratio


The observations made earlier with the
idealized wavelets, remain qualitatively valid in
case of actual records: as the ratio ac1/aH
increases, sliding displacement decreases.
However, with actual accelerograms slippage
does not always increase in inverse proportion
with ac1/aH (Fig. 13). This is explained by the
fact that, near-fault ground motions include a
large number of pulses. For instance, Figure 14
portrays the asymmetric response triggered by
the JMA record; an excitation of numerous
acceleration cycles. No uphill sliding occurred.
At each cycle a finite slippage is induced,
leading
to
successively
accumulating
displacement. However, almost all the records
include pulses in which sliding takes place
when ac1/aH has small values, but as the ratio
ac1/aH increases no sliding occurs.
Usually, a record is qualified by its peak
ground acceleration (PGA), presuming that the
magnitude of PGA is an important index of
destructiveness. However, this is not the rule for
elastoplastic systems, especially those involving
asymmetric sliding. An impressive evidence of

Influence of excitation direction


As obtained with idealized wavelets,
reversal of excitation direction may cause
significant magnification of sliding displacement.
Similar is the trend with actual accelerograms
(Figs 21-23).
The dominant direction of a record is of
importance because it determines if the
deleterious pulses of the record tend to move

82

the block uphill or downhill. Thus, Figure 21


portrays for two excitation directions (+ and -)
the response of a sliding block with ac1/aH = 0.1
on a plane of = 25, subjected to the Imperial
Valley No4 record. In the right handside of
Figure 21, notice that the distinctive long-period
pulse of T 3 sec which starts at 6.2 sec is
responsible for the slippage of 1.7 m. On the
contrary, when the record is reverted (see lefthandside of figure), this distinctive acceleration
pulse is the one who stops the block downhill
sliding, just at 0.9 m.
In fact, it is evident that all accelerograms
are recorded with a specific direction which
can not be changed. Hence which is the
purpose of applying a record with reverted
direction? Actually the direction of a record is
irreversible, but the direction of an inclined
plane is not. To appreciate the practical
significance of the change in sign of the
excitation, consider the case of the two steep
banks of a river; they have opposite slope
directions and in case of an earthquake they
subjected to the same input motion. This is
equal to the case of two identical slopes with
the same direction, which are excited by
opposite direction records. Even if they were
mathematically identical, their damage during
Chi-chi might be significantly different (see Fig.
22).

For an excitation, features such as the exact


sequence of cycles, the duration and magnitude
of acceleration pulses, and the velocity steps,
are all of major importance as they may play a
decisive role in the induced slippage,
sometimes equal important role as the
dominant frequency and the intensity of motion.
According to their form and details,
accelerograms can be grouped into one or
more of several categories:
Highfrequency records: Santa Monica,
UCLA, Lucerne, Sakarya

Records affected by forward or backward


directivity:
Fukikai,
JMA,
Takatori,
Takarazuka, TCU, Imperial Valley, Newhall,
Rinaldi, Jensen.

Considering the results of the above records


in asymmetric sliding, the following trends are
noticed:
Particularly disastrous are the records which
are characterized by a dominant velocity pulse
not only of long duration and large velocity step,
but mainly of monotonically increasing or
decreasing branch (Fig.24). By the term
monotonically increasing branch we mean that
the ratio of the velocity fluctuations amplitude
divided to the velocity step is small (Fig.24). A
necessity, to succeed this monotonic behavior
of the increasing and decreasing branch is
these variations to rise from distinctive
acceleration pulses; a forward rupture directivity
characteristic.
As can be seen in Figure 25, the velocity
pulse (of 3 sec duration and 1.5 m/s velocity
step) in the Yarimca record results in a slippage
of 4.5 m, generated by two dominant
successive acceleration pulses: the first of them
begins at the moment t 6 sec and finishes at t
7.5 sec, while the second one starts
immediately after the end of the first pulse and
finishes at t 10.5 sec. As a result, the induced
velocity pulse has a monotonically increasing
branch succeeded by a decreasing one.
On the other hand, even though the Sakarya
record includes a substantial velocity pulse of
4.5 sec duration and 0.9 m/s velocity step, it
causes a sliding displacement of only 1 m. This
may be explained by the fact that the velocity
pulse is built up by continuously fluctuating
acceleration spikes, each of which of high
magnitude.

Influence of excitation form

Records with fling-steps: Yarimca 060,


Sakarya, TCU 068 NS & EW, Lucerne

COMPARISON WITH AVAILABLE CHARTS

It is of great practical interest to compare the


results of our study with the relevant charts for
sliding displacement published by Makdisi &
Seed (1978), Sarma & Ambraseys (1967,
1978), and Yegian et al (1988). The
comparisons are displayed in Figures 26 28,
from which the following observations are
made:

Records with a single dominant velocity


pulse: Imperial Valley No 4-230 & 6-230,
Yarimca 060, Sakarya, TCU 068 NS & EW,
Lucerne 275, Rinaldi 228, Pacoima Dam
164.
Records with multiple dominant velocity
pulses: JMA, Takatori, ImperialValley No 4140, TCU 102 NS, Jensen 022.

83

1)

2)

3)

REFERENCES AND BIBLIOGRAPHY

Both in case of idealized wavelets and


actual records, the proposed curves by
Yegian et al (1988) are in very good
agreement with our results.
The curves of Makdisi & Seed (1978), and
Ambraseys & Sarma (1967, 1978)
generally lead to satisfactory results for all
idealized wavelets, but the sinus pulse.
With the latter excitation their charts
would
under-estimate
the
induced
slippage (computed here).
For most of the records (particularly the
strong near fault ground motions such as
TCU 068 NS, TCU 068 EW, TCU 102
NS, Fukiai, JMA, Takatori, and
Takarazuka) the curves of Makdisi &
Seed and Ambraseys & Sarma,
understandably, fail to predict even the
order of magnitude of the true sliding (Fig.
28).

Ambraseys, N.N., & Sarma, S.K. (1967) The


response of earth dams to strong earthquakes,
Gotechnique, Vol.17, pp 181 - 213.
Amirbekian, R.V. & Bolt, B.A. (1998) Spectral
comparison of vertical and horizontal seismic
strong ground motions in alluvial basins,
Earthquake Spectra, EERI, Vol. 14, No 4, pp
573- 595.
Baker, R. & Klein, Y. (2004) An integrated limiting
equilibrium approach for design of reinforced soil
retaining structures: Part I formulation,
Geotextiles and Geomembranes, Vol. 22, pp 151
177.
Baker, R. & Klein, Y. (2004) An integrated limiting
equilibrium approach for design of reinforced soil
retaining structures: Part II design examples,
Geotextiles and Geomembranes, Vol. 22, pp 119
150.
Bathe, K. J., & Mijailovich, S. (1988) Finite element
analysis of frictional contact problems, Journal
of Theoretical and Applied Mechanics, Vol. 7, pp
31 - 44.

CONCLUSIONS

Response in asymmetric sliding is


depending on several factors, which make the
extraction
of
specific
rules
difficult.
Nevertheless, the following general trends were
brought to light:
Peak ground acceleration does not have so
great importance in asymmetric sliding, as
do the frequency and direction of excitation.
The details, the sequence of pulses and the
form of the input motion have a dominant
influence on sliding response because of
the non-linearity of the problem.
On the basis of induced slippage, the most
devastating ground motions are these which
have a strong pulse character both in
acceleration and velocity. Such are the
records affected by forward directivity
effects.
Finally, comparison between the numerical
results of the present paper and literature,
results to the conclusion that curves of
Yegian et al describe even the extreme
reality accurately enough. On the other
hand, the proposed curves of Makdisi,
Seed,
Ambraseys
and
Sarma
underestimate, sometimes dramatically, the
slippage caused by near-fault ground
motions.

Bolt, B.A. (2004) Earthquakes, W.H. Freeman and


Company, Fifth edition, 41 Madison Avenue,
New York, NY 10010.
Bozorgnia, Y. & Bertero, V.V. (2004) Earthquake
Engineering: From Engineering Seismology to
Performance - Based Engineering, CRC Press,
2004.
Cai, Z. & Bathurst, R.J. (1996) Seismicinduced
permanent displacement of geosyntheticreinforced segmental retaining walls, Canadian
Geotechnical Journal, Vol. 33, pp 937 955.
Constantinou, M.C., & Gazetas, G. (1984)
Probabilistic seismic sliding deformations of
earth dams and slopes, Proceedings of
Specialty Conference on Probabilistic Methods
and Structural Reliability, ASCE, New York, pp
318 - 321.
Duncan, J.M. (1996) State of the art: Limit
equilibrium and finite element analysis of slopes,
Journal of Geotechnical Engineering, Vol. 122,
pp 577 596.
Elgamal, A.W.M., Scott, R.F., Succarieh, M.F., &
Yan, L. (1990) La Villita dam response during
five
earthquakes
including
permanent
deformation,
Journal
of
Geotechnical
Engineering, Vol. 116, pp 1443 - 1462.
Elms, D.G. (2000) Refinements to the Newmark
sliding block model, Proceedings of 12th World
Conference in Earthquake Engineering, pp 2132.

84

Gazetas, G. & Uddin, N. (1994) Permanent


deformation on pre-existing sliding surfaces in
dams, Journal of Geotechnical Engineering,
Vol.120, No.11, pp 2041 2061.

Michalowski, R. & You, L. (2000) Displacements of


reinforced slopes subjected to seismic loads,
Journal of Geotechnical and Geoenvironmental
Engineering, Vol. 126, No 8.

Junwu, D., Tong, M., Lee, G.C., Xiaorhai, Q. &


Wenting, B. (2004) Dynamic responses under
the excitation of pulse sequences, Earthquake
Engineering and Engineering Vibration, IEM and
MCEER, Vol. 3, No 2, pp 157 - 169.

Nadim. F., & Whitman, R.V. (1983) Seismically


induced movements of retaining walls, Journal
of Geotechnical Engineering Division, Vol. 109(7),
pp 915 - 931.
Newmark, N.M. (1965) Effect of earthquakes on
dams and embankments, Gotechnique, Vol. 15,
pp 139 -160.

Kramer, S. L. & Paulsen, S. (2002) A numerical


model for estimating seismic displacements of
reinforced steep slopes, M.S thesis, University
of Washington.

Richards, R., Elms, D.G. & Budhu, M. (1982)


Seismic bearing capacity and settlement of
foundations,
Journal
of
Geotechnical
Engineering Division.

Kramer, S. L. & Smith, M. (1997) Modified Newmark


model for seismic displacements of compliant
slopes,
Journal
of
Geotechnical
and
Geoenvironmental Engineering, Vol. 123, pp 635
644.

Richards, R. & Elms, D.G. (1979) Seismic


behaviour of gravity retaining walls, Journal of
Geotechnical Engineering Division.

Kramer, S. L. (1996) Geotechnical Earthquake


Engineering, Prentice Hall, Upper Saddle River,
NJ 07458.

Sarma, S.K. (1975) Seismic stability of earth dams


and embankments, Gotechnique, Vol. 25, No 4,
pp 743 - 761.

Lin, J.S., & Whitman, R.V. (1986) Earthquake


induced displacements of sliding block, Journal
of Geotechnical Engineering, Vol. 105, pp 1427 1434.

Seed, H.B., & Martin, G.R. (1966) The seismic


coefficient in earth dam design, Journal of the
Soil Mechanics and Foundation Division, ASCE,
Vol. 92, pp 25 - 58.

Lin, J.S. (1982) Probabilistic evaluation of the


seismically - induced permanent displacements
in earth dams, Research Report No R82-21,
Department of Civil Engineering, MIT.

Succarieh, M. (1990) Analysis of earthquake


induced large displacements in earth structures,
PhD thesis, Rensselaer Polytechnic Institute,
Troy, N.Y.

Ling, H. (2001) Recent applications of sliding block


theory to geotechnical design, Soil Dynamics
and Earthquake Engineering, Vol. 21, pp 189
197.

Wartman, J., Bray, J.D. & Seed, R.B. (2003)


Inclined plane studies of the Newmark sliding
block procedure, Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 129, pp 673
684.

Ling, H., Leshchinsky, D. & Chou, N. (2001) Post


earthquake
investigation
on
several
geosynthetic- reinforced soil retaining walls and
slopes during the Ji-Ji earthquake in Taiwan,
Soil Dynamics and Earthquake engineering, Vol.
21, pp 297 313.

Westermo, B., & Udwadia, F. (1983) Periodic


response of a sliding oscillator system to
harmonic excitation, Earthquake Engineering
and Structural Dynamics, Vol. 11, pp 135 -146.
Yegian, M.K., Marciano E.A., & Ghahraman, V.G.
(1991)
Earthquake
induced
permanent
deformations: a probabilistic approach, Journal
of Geotechnical Engineering, Vol. 117, pp 35 50.

Makdisi, F.I. & Seed, H.B. (1978) Simplified


procedure for estimating dam and embankment
earthquake induced deformations, Journal of
Geotechnical Engineering Division, ASCE, Vol.
104, pp 849 867.

Yegian, M.K., Marciano E.A., & Ghahraman, V.G.


(1988) Integrated seismic risk analysis for earth
dams, Report No CE-88-15.

Makris, N. (1989) Analysis of harmonically excited


sliding systems, M.S thesis, State University of
New York, Buffalo, N.Y.

85

Fk

d(t)

d(t)

a(t)

a H(t)

(a)
Fig 1:

(b)

Schematic representation of Newmarks model used in the parametric study. (a) A


rigid block resting on an inclined sliding interface governed by Coulombs friction
law, (b) forces acting on a rigid block that tends to slide downwards when subjected
to upward excitation.

1.5

1.2
=0
=5

= 25

:m

0.9

0.6

0.3

0
0

0.1

0.2

0.3

0.4

0.5

0.6

ac1 / aH
Fig 2:

Sliding displacement according to acceleration ratio ac1 / aH in case of a Ricker


excitation with frequency fH = 1 Hz and peak acceleration aH = 10 m/s2. Notice the
difference between symmetric ( = 0o) and asymmetric sliding ( 0o).

86

1 .5

25

o
1.5
f = 1fH=
z

1 .2

f = 1fH=
z

1 Hz

f = 2f H=z 2

Hz

f = 4f H=z 4

Hz

H
H

0 .9

: m

1.2

0 .6

0 .3

0 .3

Hz

f = 4f H=z 4

Hz

H
H

0 .9

0 .6

1 Hz

f = 2f H=z 2

0
0

0 .1

0 .2

0 .3

0 .4

0 .5

0 .6

0 .1

0 .2

0 .3

a c1 / a

Fig 3:

0 .4

0 .5

0 .6

a c1 / a

Asymmetric sliding spectra illustrating the influence of critical acceleration ratio ac1 /
aH , for different excitation frequencies of a Ricker wavelet with maximum
acceleration aH = 10 m/s2. It is clear that by increasing the inclination of the plane the
slippage is also increasing.

Ground

Sliding block

o
12

-6

-6

acceleration : m / s

12

-1 2

-1 2

velocity : m / s

-0 .6

-0 .6

-1 .2

-1 .2

-1 .8

-1 .8

-2 .4

-2 .4

5
0 .8

0 .8

0 .4
d(t) : m

0 .4

0.76 m

0.40 m

-0 .4

-0 .4
0

t : s

Fig 4:

t : s

Effect of acceleration ratio to sliding response induced by a sinusoidal pulse of fH =


1.5 Hz. The acceleration ratio is ac1/ a H = 0.05 for the plots in the left handside and
0.1 for those in the right. The shaded areas in the velocity time histories represent the
downward (dotted) or upward (striped) sliding.

87

Ground
Sliding block

acceleration : m / s 2

Frequency: fH

= 1 Hz

Frequency:

12

12

velocity : m / s

0.5

1.5

2.5

3.5

4.5

0.5

2.5

2.5

1.5

1.5

0.5

0.5

1.5

2.5

3.5

1.5

2.5

3.5

4.5

0
0.5

1.5

2.5

3.5

0.5

4.5

4.5

1.86 m

1.5
d(t) : m

fH = 2 Hz

1.5

0.5

0.5

0.47 m

0.5

1.5

2.5

3.5

4.5

0.5

t:s
Fig 5:

1.5

2.5

3.5

4.5

t:s

Acceleration, velocity and displacement time-histories induced by a T-Ricker pulse of


two different frequencies (inclination angle = 5o and ac1/aH = 0.1 ). The single
dominant slip (dotted area) is the outcome of the one-sided acceleration pulse which
consequently generates a monotonically increasing velocity time history.

88

F re q u e n c y: f H = 1 .5 H z

F re q u e n c y: f H = 3 H z

12

12

-6

-6

acceleration : m / s

G ro u n d

-1 2

-1 2

velocity : m / s

S lid in g b lo c k

-0 .8

-0 .8

-1 .6

-1 .6

-2 .4

-2 .4
0

d(t) : m

4.25 m
4

1.10 m

t:s

Fig 6:

t:s

Acceleration, velocity and displacement time-histories for a sinusoidal excitation of


one cycle (inclination angle = 25o and acp1/aH = 0.05 ). The response reach the 4.25
m when the frequency of the pulse is fH = 1.5 Hz (left) whereas for the higher
frequency of 3 Hz (right) slippage falls to 1.10 m.

1g

25o
4.5

10

fH = 2 Hz

3.6

aH / VH2

fH = 1 Hz

: m

fH = 2 Hz

2.7

1.8

0.9

fH = 1 Hz

0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Fig 7:

0.1

0.2

0.3

0.4

0.5

0.6

0.7

ac1 / a

ac1 / a

Dimensional (left) and normalized (right) sliding spectra for a T-Ricker wavelet in
case of a 25o inclined plane.

89

1g

1g

25

1.5

3.2

fH = 0.67 Hz

fH = 0.67 Hz
1.2

fH = 1.5 Hz

2.4

aH / VH2

1.6

0.8

fH = 3 Hz
0.9

0.6

0.3

0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.1

0.2

ac1 / a

Fig 8:

= 5

0.5

0.6

= 25

10

acceleration : m / s

0.4

G ro u n d
5

G ro u n d
5

S lid in g b lo c k

-5

-5

-1 0

S lid in g b lo c k

-1 0
0

velocity : m / s

0.3

ac1 / a

Normalization effect of sliding spectra in case of a sinusoidal wavelet of one cycle.


Observe that dimensional slippage (left figure) is varying for every frequency fH,
however the normalized response (right plot) converge to a single line.

10

0 .5

0 .5

-0 .5

- 0 .5

-1

-1

-1 .5

- 1 .5
0

d(t) : m

: m

fH = 3 Hz

fH = 1.5 Hz

1 .5

1 .5

1 .2

1 .2

0 .9

0 .9

0 .6

1.47 m

0 .6

0.35 m

0 .3

0 .3
0

- 0 .3

- 0 .3
0

t : s

t : s

Fig 10: Effect of slope inclination to sliding response induced by a Tsang wavelet with fH = 1
Hz. The acceleration ratio is ac1/ a H = 0.05. The shaded areas in the velocity time
histories represent the downward (dotted) or upward (striped) sliding.

90

acceleration : m / s

Frequency: fH

= 1 Hz

Frequency:

10

10

-5

-5

Ground
Sliding block

-10

-10
0

velocity : m / s

fH = 2 Hz

1.2

1.2

0.6

0.6

-0.6

-0.6

-1.2

-1.2

0.4

0.4

0.38 m
0.2

0.10 m

d(t) : m

0.2

-0.2

-0.2

t:s

t:s

Fig 11: Response time-histories for Ricker excitation of frequency fH = 1 Hz (left) and fH = 2
Hz (right) in case of a plane with inclination = 5o and ac1/ a H = 0.1. The pulse is
imposed inverted to study sliding with respect to triggering input direction.

91

Ground
Sliding block

acceleration : m / s 2

25o
12

12

-6

-6
-12

-12

velocity : m / s

2.4

2.4

1.2

1.2

-1.2

-1.2

-2.4

-2.4
0

d(t) : m

4.5

4.5

3.6

3.6

2.7

2.7

1.8

4.25 m

1.8

0.62 m

0.9

0.9

0
0

t:s

t:s

Fig 12: By reversing the polarity of an one cycle sinusoidal pulse (fH = 1.5 Hz) the resulting
slippage, in case of an inclination angle = 25o and ac1/aH = 0.05, is increased seven
times (from 0.62 m to 4.26 m).

92

0.81 g
Ground
Sliding block
o

25

0.5

2.71 m

2
d(t) : m

velocity : m / s

1
-0.5
0

-1
0

10

12

14

10

12

14

0.5
0

0.85 m

1
-0.5
0

-1
0

10

12

14

10

12

14

0.5
2
d(t) : m

velocity : m / s

d(t) : m

velocity : m / s

-0.5

0.27 m
0

-1
0

10

12

14

t:s

10

12

t:s

Fig 13: Influence of acceleration ratio ac1/ a H on asymmetric sliding for the inverted JMAi
record (NS component and inclination angle = 25o). The two plots of the first top
line correspond to a ratio ac1/ a H = 0.05, whereas the plots of the middle and bottom
line correspond to critical acceleration ratios of 0.2 and 0.4, respectively.

93

14

Ground
Sliding block

25
6

acceleration : m / s

3
0
-3
-6
-9

10

12

14

10

12

14

velocity : m / s

0 .5
0
-0 .5
-1

d(t) : m

2.58 m

10

12

14

Tim e : sec

Fig 14:

Acceleration and velocity time-histories of a rigid block (black colored lines) and the input base
excitation (grey colored lines), as well as slippage time-history of the sliding block induced by
the JMA ground motion (NS component); in case of a slope with inclination angle =25o and
critical acceleration ratio ac1/aH = 0.05. The slippage of 2.58 m at the end of the motion is the
result of all the previously accumulated slides.

acceleration : m / s

4
2
0
-2

0.34 g

-4
0

12

15

18

21

12

15

18

21

velocity : m / s

2.97 m/s
2

-2
0

d(t) : m

12

8.81 m

12

15

18

21

T im e : s e c

Fig 15:

Asymmetric response time-histories for the TCU 068 NS ground motion ( = 25o and ac1/ a H =
0.05 ). The 8.8 m of total slippage is induced by a large sequence of remarkably long duration
acceleration pulses. Observe the outstanding ground velocity pulse of almost 6 sec duration
and 3 m/s magnitude.

94

Ground
Sliding block

25
4

-2

-2

-2

-4

-4

acceleration : m / s

velocity : m / s

12

18

24

12

18

24

-2

-2
0

d(t) : m

-4
0

12

18

20

20

15

15

8.8

12

18

24

13.6

10

12

18

24

12

18

24

16.0

0
0

24

18

15

10

12

20

10

-2
0

24

12

t : s

18

24

12

t : s

18

24

t : s

Fig 16: Simple approximation of the TCU 068- NS record through a series of rectangular
pulses (inclination angle = 25o and ac1/aH = 0.05 ). The left column plots show the
response to the detailed ground motion, while the middle and the right column display
both the simplified acceleration excitations that were used. Even though the first
major slippage of nearly 6.5 m (dotted line) is approximated by the simplified motions,
the cut off of the details leads to substantial increase in the final result.
4

-2

-2

-2

-4

-4

acceleration : m / s

velocity : m / s

18

27

-4
0

18

27

-2

-2

-2

-4
9

18

27

30

18

27

18

27

-4

-4

18

27
30

30

d(t) : m

23.3
20

25.3

10

27.1

20

10

10

0
0

18

t : s

27

20

18

t : s

27

18

t : s

Fig 17: Simple simulation of the inverted TCU 068-NS record through a series of rectangular
pulses (inclination angle = 25o and ac1/aH = 0.05 ). The final displacements in all
three cases (the actual record and the approximations) are in reasonable agreement.

95

27

Ground
Sliding block

acceleration : m / s 2

5o
10

10

-5

-5

velocity : m / s

-10

-10
2

10

12

14

1.5

1.5

0.5

0.5

-0.5

-0.5

-1

-1

10

12

14

10

12

14

-1.5

-1.5
0

d(t) : m

25

10

12

14

0.8

0.8

0.6

0.6

0.4

0.4

0.53 m

0.2

0.78 m

0.2
0

-0.2

-0.2

10

12

14

t:s

10

12

14

t:s

Fig 18: Effect of slope inclination to asymmetric sliding for the Fukiai ground motion. The
plots correspond to a ratio ac1/ a H = 0.2. In case of = 5o, the small inclination allows
an uphill sliding of the block (presented between the dotted lines in left handside
figures).

96

0.85 g

25

3.2

3.2

ac1/ a H = 0.05

d(t) : m

2.4
1.6

1.6

0.59

0.8

0.8

0
-0.8

-0.8
0

12

16

3.2

d(t) : m

12

16

3.2

ac1/ a H = 0.2

2.4

ac1/ a H = 0.2

2.4

1.6

1.6
0.31

0.8

0.78

0.8

-0.8

-0.8
0

12

16

3.2

12

16

3.2

ac1/ a H = 0.4

2.4
d(t) : m

3.13

2.4

ac1/ a H = 0.4

2.4

1.6

1.6
0.10

0.8
0

-0.8

-0.8
0

12

0.16

0.8

16

t:s

t:s

Fig 19: Influence of the symmetric (left) or asymmetric (right) nature of sliding to the
response induced by the Fukiai ground motion.

97

12

16

25

0.21 g

ac1/ a H = 0.05

d(t) : m

4
3
2

0.79

-0.1

-1

-1
0

12

16

20

12

16

20

ac1/ a H = 0.1

ac1/ a H = 0.1

4
d(t) : m

4.31

2.73

3
2

0.82

0.18

-1

-1
0

12

16

20

12

16

20

ac1/ a H = 0.2

4
d(t) : m

ac1/ a H = 0.05

ac1/ a H = 0.2

2
1

1.00

0.50

0.02

-1

-1
0

12

16

20

t:s

12

16

t:s

Fig 20: Symmetric (left) and asymmetric (right) sliding response to the Yarimca-060 record.
Observe the substantial larger slippage in case of the inclined plane, particularly for
small acceleration ratio ac1/ a H

98

20

Ground

acceleration : m / s 2

25

Sliding block

-2

-2

-4

-4

velocity : m / s

10

13

16

0.8

0.8

0.4

0.4

-0.4

-0.4

10

13

16

10

13

16

-0.8

-0.8
1

d(t) : m

10

13

16

1.5

1.5

0.93 m

1
0.5

1.70 m

1
0.5

0
1

10

13

16

t:s

10

13

t:s

Fig 21: Effect of excitation polarity on asymmetric sliding for the Imperial Valley No4 record
(230o component, inclination angle = 25o, ratio ac1/ a H = 0.1). Observe that by
reverting the triggering the subsequent two downward slides (dotted areas in left) are
replaced by a single major slide (dotted area in right).

99

16

acceleration : m / s 2

0.34 g
2

-2

-2

0.34 g

-4
0

10

-4
15

20

25

27

10

15

d(t) : m

18

25.30

18

8.81
9

0
0

10

15

20

25

10

15

20

25

27

27

ac1/ a H = 0.1
d(t) : m

25

27

ac1/ a H = 0.05

ac1/ a H = 0.1
18

18

5.46

12.90

0
0

10

15

20

25

27

10

15

20

25

27

ac1/ a H = 0.2
d(t) : m

20

ac1/ a H = 0.2

18

18

5.73

1.70
0

0
0

10

15

20

25

t:s

10

15

20

25

t:s

Fig 22: Sliding response of the TCU 068-NS record triggering a 25o inclined plane with the
peak ground acceleration of 0.34 g acting downwards (left handside) and upwards
(right handside).

100

Ground
Sliding block

acceleration : m / s 2

25
10

10

-5

-5

-10

-10

velocity : m / s

10

-1

-1

-2

10

10

-2
0

10
2.2

2.2

d(t) : m

2.00
1.04

1.1

1.1

0
0

10

t:s

t:s

Fig 23: Acceleration, velocity and displacement time-histories for the Rinaldi record (228o
component) when imposed parallel to the sliding interface (inclination angle = 25o
and ac1/aH = 0.1 ). Notice the asymmetric response of the block when the excitation is
inverted (plots in right); the well-shaped forward directivity pulse, shown between the
dotted lines, causes a major slippage of 2 m.

101

10

Ground
Sliding block

25

acceleration : m / s 2

5
2.5
0
-2.5
-5
0

12

15

18

21

12

15

18

21

1.2

velocity : m / s

Velocity Step
0.6

Fluctuation

0
-0.6
-1.2
0

d(t) : m

4
3
2

4.55

1
0
0

12

15

18

Time : sec

Fig 24: Acceleration and velocity time-histories of the rigid block and the input base
excitation, as well as slippage time-history of the sliding block triggered by the Jensen
record (22o component) for a slope of inclination =25o and critical acceleration ratio
ac1/aH = 0.05. The slippage of4.55 m is the outcome of three major sliding episodes
(dotted areas) induced by three subsequent velocity pulses resulting from the forward
directivity acceleration pulses (starting at 4 sec and ending at 12 sec).

102

21

Ground
Sliding block

25

-2

-2

acceleration : m / s

-4

-4

velocity : m / s

12

15

0.5

0.5

12

15

18

21

12

15

18

21

-0.5

-0.5
0

d(t) : m

12

15

4.31

1.01

1
0

0
0

12

15

t:s

Fig 26:

12

15

18

21

t:s

Acceleration, velocity and displacement time-histories for the Sakarya record (left)
and the Yarimca-060 component (right) of a block on an inclined plane with = 25o
and acceleration ratio acp1/aH = 0.05.

/ (aH TH )

0 .1

0 .0 1
Y e g ia n e t a l ( 1 9 8 8 )
S in u s o id a l

Sarma

tr ia n g u la r

(1976)

R ic k e r

0 .0 0 1

T - R ic k e r
s insinusoid
us
Tsang
0 .0 0 0 1
0

0 .1

0 .2

0 .3

0 .4

0 .5

0 .6

0 .7

0 .8

0 .9

a c1 / a

Fig 26:

Comparison of normalized slippage induced by idealized wavelets with the published


results of Yegian et al (1988) & Sarma (1979).

103

0.1

0.1

0.01

0.01

/ (aH TH )

JM A

TC U 068 - N S
0.001

Takatori

0.001

TC U 068 - E W

Takarazuka
Fukiai

TC U 102 - N S
0.0001

0.0001
0

0.1

0.2

0.3

0.4

0.5

0.6

0.1

0.2

0.3

0.4

0.5

0.6

a c1 / a H

/ (aH TH )

0.1

25

0.01

IV 4-140
0.001

IV 4-230

Y egian et al (1988)

IV 6-230

S inusoidal
Triangular

0.0001
0

0.1

0.2

0.3

0.4

0.5

Sarma
(1976)

0.6

a c1 / a H

: cm

Fig 27:

Comparison with the published results of Yegian et al (1988) & Sarma (1979) of the
normalized slippage induced by: Chi-chi records (top-left plot), Kobe records (topright), and Imperial Valley records (bottom-left). Each excitation imposed normally
and inverted.

1000

1000

100

100

10

10

y
JM A
Fukiai

IV 4-140
IV 4-230
IV 6-230

Takatori
Takarazuka
0.1

0.1
0

0.1

0.2

0.3

0.4

0.5

0.1

0.2

0.3

0.4

a c1 / a H
10000

: cm

1000

100

25

10

M = 6.5

T C U 068 - N S

T C U 068 - E W

M = 7.5

T C U 102 - N S
0.1
0

0.1

0.2

0.3

0.4

0.5

Makdisi & Seed


(1976)

Sarm a & Am braseys

a c1 / a H

Fig 28:

Comparison with the published results of Makdisi & Seed(1976) of the sliding
response induced by: Chi-chi records (top-left plot), Kobe records (top-right), and
Imperial Valley records (bottom-left). Each excitation imposed normally and inverted.

104

0.5

A Replacement to the MononobeOkabe Equations


by Stress Limit Analysis
G. Mylonakis, P. Kloukinas, C. Papantonopoulos
University of Patras, Greece

Abstract
A closed-form stress plasticity solution is presented for gravitational and earthquakeinduced earth pressures on retaining walls. The proposed solution is essentially an
approximate yield-line approach, based on the theory of discontinuous stress fields, and
takes into account the following parameters: (1) weight and friction angle of the soil material,
(2) wall inclination, (3) backfill inclination, (4) wall roughness, (5) surcharge at soil surface,
and (6) horizontal and vertical seismic acceleration. Both active and passive conditions are
considered by means of different inclinations of the stress characteristics in the backfill.
Results are presented in the form of dimensionless graphs and charts that elucidate the
salient features of the problem. Comparisons with established numerical solutions, such as
those of Chen and Sokolovskii, show satisfactory agreement (maximum error for active
pressures about 10%). It is shown that the solution does not perfectly satisfy equilibrium at
certain points in the medium, and hence cannot be classified in the context of limit analysis
theorems. Nevertheless, extensive comparisons with rigorous numerical results indicate that
the solution consistently overestimates active pressures and under-predicts the passive.
Accordingly, it can be viewed as an approximate lower-bound solution, than a mere
predictor of soil thrust. Compared to the Coulomb and Mononobe-Okabe equations, the
proposed solution is simpler, more accurate especially for passive pressures) and safe, as it
overestimates active pressures and underestimates the passive. Contrary to the
aforementioned solutions, the proposed solution is symmetric, as it can be expressed by a
single equation - describing both active and passive pressures - using appropriate signs for
friction angle and wall roughness.

INTRODUCTION

and need not be repeated herein [Kramer,


1996; Wood, 1973; Steedman & Zeng, 1990;
Veletsos & Younan, 1994; Theodorakopoulos et
al., 2001). Given their practical nature and
reasonable predictions of actual dynamic
pressures (e.g. Ebeling et al., 1992; Sheriff et.
al., 1982; Veletsos & Younan, 1994;
Theodorakopoulos et al., 2001), solutions of
this type are expected to continue being used
by engineers for a long time to come. This
expectation does not seem to diminish by the
advent
of
displacement-based
design
approaches, as the limit thrusts provided by the
classical methods can be used to predict the
threshold (yield) acceleration beyond which
permanent dynamic displacements start to
accumulate [Kramer, 1996; Pecker, 1995;
Richards & Elms, 1985; Whitman & Liao, 1985;

The classical equations of Coulomb [Coulomb,


1776; Heyman, 1972; Lambe, 1969; Clough &
Duncan, 1990; Chen, 1975) and MononobeOkabe (Okabe, 1924; Mononobe & Matsuo,
1929; Matsuo & Ohara, 1960; Seed & Whitman,
1970; Chen, 1975; Ebeling et al., 1992; Kramer
1996) are being widely used for determining
earth pressures due to gravitational and
earthquake loads, respectively. The MononobeOkabe solution treats earthquake loads as
pseudo-dynamic,
generated
by
uniform
acceleration in the backfill. The retained soil is
considered a perfectly plastic material, which
fails along a planar surface, thereby exerting a
limit thrust on the wall. The theoretical
limitations of such an approach are well known

105

can be criticized on the following important


aspects: (1) in the context of limit analysis, their
predictions are unsafe; (2) their accuracy (and
safety) diminishes in the case of passive
pressures on rough walls, (3) the mathematical
expressions are complicated and difficult to
verify 1 , (4) the distribution of tractions on the
wall are not predicted (typically assumed linear
with depth following Rankines solution), (5)
optimization of the failure mechanism is
required in the presence of multiple loads, to
determine a stationary (optimum) value of soil
thrust, and (6) in the context of limit-equilibrium
analysis, stress boundary conditions are not
satisfied, as the yield surface does not
generally emerge at the soil surface at the
required angles of 45o /2.
In light of the above arguments, it appears
that the development of a closed-form solution
of the stress type for assessing seismicallyinduced earth pressures would be desirable. It
will be shown that the proposed solution,
although approximate, is mathematically
simpler than the existing kinematic solutions,
offers
satisfactory
accuracy
(maximum
deviation for active pressures against rigorous
numerical solutions less than 10%), yields
results on the safe side, satisfies stress
boundary conditions, and predicts the point of
application of soil thrust. Last but least, the
solution is symmetric that is, it active and
passive pressures can be described by a single
formula, using appropriate signs for the friction
angle and wall roughness. Apart from its
intrinsic theoretical interest, the proposed
analysis can be used for the assessment and
improvement of other related methods.

Psarropoulos et. al., 2005).


Owing to the translational and staticallydetermined failure mechanisms employed, the
limit-equilibrium Mononobe-Okabe solutions
can be interpreted as kinematic solutions of limit
analysis (Collins, 1973). The latter solutions are
based on kinematically admissible failure
mechanisms in conjunction with a yield criterion
and a flow rule for the soil material, both of
which are enforced along pre-specified failure
surfaces (Chen, 1975; Pecker, 1995; Caquot &
Kerisel, 1948 ; Finn, 1967; Salencon, 1974).
Stresses outside the failure surfaces are not
examined and, thereby, equilibrium in the
medium is generally not satisfied. In the realm
of associative and convex materials, solutions
of this type are inherently unsafe that is, they
underestimate
active
pressures
and
overestimate the passive (Chen, 1975; Finn,
1967; Davis & Selvadurai, 2002; Salencon,
1974).
A second group of limit-analysis methods,
the stress solutions, make use of pertinent
stress fields that satisfy the equilibrium
equations and the stress boundary conditions,
without violating the failure criterion anywhere in
the medium (Davis & Selvadurai, 2002;
Atkinson, 1981; Parry, 1995). On the other
hand, the kinematics of the problem is not
examined and, therefore, compatibility of
deformations is generally not satisfied. For
convex materials, formulations of this type are
inherently safe that is, they overestimate active
pressures and underestimate the passive
(Chen, 1975; Davis & Selvadurai, 2002;
Atkinson, 1981). The best known such solution
is that of Rankine, the applicability of which is
severely restricted by the assumptions of
horizontal backfill, vertical wall and smooth soilwall interface. In addition, the solution may be
applied only if the surface surcharge is uniform
or non-existing. Owing to difficulties in deriving
pertinent stress fields for general geometries,
the vast majority of limit-analysis solutions in
geotechnical design are of the kinematic type
(Seed & Whitman, 1970; Ebeling et al., 1992;
Chen, 1975; Kramer 1996, Atkinson, 1981]. To
the best of the authors knowledge, no simple
closed-form solution of the stress type has been
derived for seismic earth pressures.
Notwithstanding the theoretical significance
and practical appeal of the Coulomb and
Mononobe-Okabe solutions, these formulations

METHODOLOGY

The problem under investigation is depicted


in Figure 1: a slope of dry cohesionless soil
retained by an inclined gravity wall, is subjected
to plane strain conditions under the combined
action of gravity (g) and seismic body forces (ah
x g) and (av x g) in the horizontal and vertical
direction, respectively. The problem parameters
are: height (H) and inclination () of the wall,
1

The story of a typographical error in the MononobeOkabe formula that appeared in a seminal article of the
early 1970s and subsequently propagated in a large
portion of the literature, is indicative of the difficulty in
checking the mathematics of these expressions (Davies et
al [41]).

106

inclination () of the backfill; roughness () of


the wall-soil interface; friction angle () and unit
weight () of the soil material, and surface
surcharge (q). Since backfills typically consist of
granular materials, cohesion in the soil and
cohesion at the soil-wall interface are not
studied here.

tan e =

inclined
backfill

+ e
cohesionless soil
( )

backfill, under zero rotation. Both assumptions


have important implications in the distribution of
earth pressures on the wall, as explained
below.
The resultant body force in the soil is acting
under an angle e from vertical towards the
backfill,

+a
+a

inclined wall,
roughness ()

Fig 1: The problem under consideration

ah
1 av

(1)

which minimizes passive resistance. In


accordance with the rest of the literature,
positive av is upward (downward ground
acceleration). However, its influence on earth
pressures, although included in the analysis, is
not studied numerically here, as it is usually
minor and often neglected in design (Ebeling et
al., 1992; Whitman & Liao, 1985).
In the absence of surcharge, the MononobeOkabe solution to the above problem is given
by the well-known formula (Kramer, 1996):

In addition, since the vibrational characteristics


of the soil are neglected, the seismic force is
assumed to be uniform in the backfill. Also, the
wall can translate away from, or towards to, the

2P
K E = E2 =
H

cos2 ( e + )

sin ( + ) sin ( e + )

cos e cos2 cos ( e + ) 1


cos ( e + ) cos ( )

e <

where PE denotes the limit seismic thrust on the


wall (units = F/L) and KE is the corresponding
earth pressure coefficient. In the above
representation (and hereafter), the upper sign
refers to active conditions (PE = PAE, KE = KAE),
and the lower sign to passive (PE = PPE, K =
KPE).
A drawback of the above equation lies in the
difficulty in interpreting the physical meaning especially signs - of the various terms (Davis &
Selvadurai, 2002, footnote in p4). As will be
shown below, the proposed solution is free of
this problem.
To prevent slope failure when inertial action
is pointing towards the wall, the seismic angle
e should not exceed the difference between
the friction angle and the slope inclination.
Therefore, the following constraint applies
(Ebeling et al., 1992):

(2)

(3)

A similar relation can be written for the case


where inertial action is pointing towards the
backfill, but it is of limited practical interest and
will not be discussed here.
To analyze the problem, the backfill is
divided into two main regions subjected to
different stress fields, as shown in Figure 2: the
first region (zone A) is located close to the soil
surface, whereas the second (zone B) close to
the wall. In both regions the soil is assumed to
be in a condition of impeding yielding under the
combined action of gravity and earthquake
body forces. The same assumption is adopted
for the soil-wall interface, which, however, is
subjected exclusively to contact stresses. A
transition zone between regions A and B is
introduced later on.
Fundamental to the proposed analysis is the
assumption that stresses close to the soil

107

which are valid for static conditions (h = v = 0)


and satisfy the stress boundary conditions at
the surface. Equations (4) suggest that the ratio
of shear to normal stresses is constant (tan) at
all depths, and that points at the same depth
are subjected to equal stresses. Note that due
to static determinacy and anti-symmetry, the
above relations are independent of material
properties and asymptotically exact at large
distances from the wall.
Considering the material to be in a condition
of impeding yielding, the Mohr circle of stresses
in region A is depicted in Figure 3. The different
locations of the stress point (, ) for active
and passive conditions and the different
inclinations of the major principal plane
(indicated by heavy lines) are apparent in the
graph.

surface can be well approximated by those in


an infinite slope, as shown in Figure 2. In this
region (A), the inclined soil element shown is
subjected to canceling actions along its vertical
sides. Thus equilibrium is achieved solely under
body forces and contact stress acting at its
bottom face.
Based
on
this
physically-motivated
hypothesis, the stresses and at the base
of the inclined element are determined from the
following expressions:

= z +
cos 2
cos

(4a)

= z +
sin cos
cos

(4b)

ZONE A

( , )

( , )

passive
case

unit length

active
case

rface
soil su

q
soil
surface

1
1

1A

1 +
SA

active

passive

ZONE B

(w , w )

ZONE A

passive

wall
plane

(w , w )

2+

passive

passive

1B

SB

active

(w , w )
active

wall
plane

(w , w )

Fig 3: Mohr circles of effective stresses and


inclination of the major principal planes in
zones A and B under gravitational loading.

active

ZONE B

From the geometry of Figure 3, the normal


stress is related to mean stress SA through
the proportionality relation

wall length
L = H / cos

Fig 2: Stress fields close to soil surface (Zone A)


and the wall (Zone B)

108

= S A [1 sin cos( 1 )]

transition in the orientation of principal planes in


the two zones, a logarithmic stress fan 2 is
adopted in this study, centered at the top of the
wall. In the interior of the fan, principal stresses
are gradually rotated by the angle separating
the major principal planes in the two regions, as
shown in Figure 4.

(5)

where 1 denotes the Caquot angle [23, 28]


given by

sin 1 =

sin
sin

(6)

For points in region B, it is assumed that


stresses are functions exclusively of the vertical
coordinate and obey the strength criterion of the
frictional soil-wall interface, as shown in Figure
2. Accordingly, at orientations inclined at an
angle from vertical,

w = w tan

AB

zone B

(7)

ACTIVE CONDITIONS

1
2
2
2 +
2

PASSIVE CONDITIONS

AB
z

(8)

zone B

where 2 is the corresponding Caquot angle


given by,

sin
sin 2 =
sin

zone A

2
2
2

where w and w are the normal and shear


tractions on the wall, at depth z. The above
equation is asymptotically exact for points in the
vicinity of the wall. The corresponding Mohr
circle of stresses is depicted in Figure 3. The
different signs of shear tractions for active and
passive conditions follow the directions shown
in Fig 2 (passive wall tractions pointing upward,
active tractions pointing downward), which
comply with the kinematics of the problem. This
is in contrast with the widespread view that
solutions based on equilibrium totally ignore the
displacement field (Papantonopoulos & Ladanyi
1973). From the geometry of Figure 3, the
normal traction w is related to the mean stress
SB through the expression

w = S B 1 sin cos ( 2 )

1 +
2

zone A

Fig 4: Rotation of major principal planes between


zones A and B for active and passive
conditions

(9)

In light of the foregoing, it becomes evident


that the orientation of stress characteristics in
the two regions is different and varies for active
and passive conditions. In addition, the mean
stresses SA and SB generally do not coincide,
which suggests that a Rankine-type solution
based on a single stress field is not possible.
To determine the separation of mean
stresses SA and SB and ensure a smooth

This additional condition is written as (Chen


1975):

S B = S A exp ( 2 tan )
2

(10)

This should not be confused with log-spiral shaped


failure surfaces used in kinematic solutions of related
problems.

109

The negative sign in the above equation


pertains to the case where SB < SA (e.g., active
case) and vice versa. The above equation is an
exact solution of the governing Ktter equations
for a weightless frictional material. For a
material with weight, the solution is only
approximate as Kotters equations are not
perfectly satisfied (Davis & Selvadurai, 2002;
Atkinson, 1981; Parry, 1995). In other words,
the log spiral fan accurately transmits stresses
applied at its boundaries, but handles only
approximately body forces imposed within its
volume. The error is expected to be small for
active conditions (which are of key importance
in design), because of the small opening angle
of the fan, and bigger for passive conditions. As
a result, the above solution cannot be
interpreted in the context of limit analysis
theorems. Nevertheless, it will be shown that
these violations are of minor importance from a
practical viewpoint.

is twice the angle separating the major principal


planes in zones A and B (Figure 4). The
convention regarding double signs in the above
equations is as before.
It is also straightforward to show that the
surcharge coefficient Kq is related to K through
the simple expression

K q = K

The total thrust on the wall due to surcharge


and gravity loading is obtained by the wellknown expression (Chen 1975):
(11)

which is reminiscent (though not equivalent) of


the bearing capacity equation of a strip surface
footing on cohesionless soil. In the above
equation, Kq and K denote the earth pressure
coefficients due to surcharge and self-weight,
respectively.
Combining Eqns (5), (8) and (10), and
integrating over the height of the wall, it is
straightforward to show that the earth pressure
coefficient K is given by [39]

K =

cos ( ) cos
cos cos 2

Solution including Earthquake Loading

Recognizing that earthquake action imposes


a resultant thrust in the backfill inclined by a
constant angle e from the vertical (Fig 1), it
becomes apparent that the pseudo-dynamic
problem does not differ fundamentally from the
corresponding static problem, as the former can
be obtained from the latter through a rotation of
the reference axes by the seismic angle e, as
shown in Fig 5. In other words, considering e
does not add an extra physical parameter to the
problem, but simply alters the values of the
other variables. This property of similarity was
apparently first employed by Briske (1927) and
later by Arango (Seed & Whitman, 1970;
Ebeling et al., 1992) in the analysis of related
problems.

(12)

1 sin cos ( 2 )

exp( 2 tan )
1 sin cos ( 1 )
where

2 = 2 ( 1 + ) + 2

(14)

which coincides with the kinematic solution of


Chen & Liu (1990), established using a
Coulomb mechanism. Note that for a horizontal
backfill ( = 0), coefficients Kq and K coincide
regardless of wall inclination and material
properties. Equation (14) represents an exact
solution for a weightless material with
surcharge. A simplified version of the above
solutions, restricted to the special case of a
vertical wall with horizontal backfill and no
surcharge (e = 0; = = 0; q =0), has been
derived by Lancelotta (2002). Another simplified
solution, which, however, contains some
algebraic mistakes (see application example at
the end of the paper) and is restricted to active
conditions and no surcharge, has been
presented by Powrie (1997).

Solution without Earthquake Loading

1
P = K q q H + K H 2
2

cos
cos ( )

(13)

110

strength parameters and are invariant to the


transformation.
In the light of the above developments, the
soil thrust including earthquake action can be
determined from the modified expression:

1
PE = K q* q* H * + K * * H *2
2

H*

in which parameters , , H, , and q have


been
replaced
by
their
transformed
counterparts. The symbols Kq* and K* denote
the surcharge and self-weight coefficients in the
modified geometry, respectively.
Substituting Eqns (15) through (19) in Eqn
(20) yields the modified earth pressure
coefficient

Fig 5: Similarity transformation based on a rotation


of the reference axes, for analyzing the
pseudo-dynamic seismic excitation as a
gravitational problem. Note the different wall
height, backfill slope, and wall inclination in
the modified geometry. Note also that the
rotation should be performed in the opposite
(clockwise) direction for passive pressures
(e<0)

KE = (1 av )

(15)

*= + e

(16)

H * = H cos ( + e ) / cos

(17)

* = (1 av ) / cos e

(18)

q * = q (1 av ) / cos e

(19)

cos ( ) cos( + e )
cos e cos cos2

(21)

1 sin cos ( 2 )

exp( 2 E tan )
*
1 sin cos 1 ( + e )

which encompasses seismic action and can be


used in the context of Eqn (11). In the above
equation,

Application of the concept to the present


analysis yields the following algebraic
transformations, according to the notation of
Figure 5:

*= + e

(20)

2 E = 2 ( 1* + ) + 2 e

(22)

is twice the revolution angle of principal


stresses in the two regions under seismic
conditions; 1* is equal to Arcsin[sin(+e)
/sin], following Eqns (6) and (15).
The seismic earth pressure coefficient KEq is
obtained as

K Eq = K E

The modification in and q is due to the


change in length of the corresponding vectors
(Fig 1) as a result of inertial action. To obtain
Eqn (19), it has been tacitly assumed that the
surcharge responds to the earthquake motion in
the same manner as the backfill and, thereby,
the transformed surcharge remains vertical.
Note that this is not an essential hypothesis -
just a convenient (reasonable) assumption from
an analysis viewpoint. Understandably, the

cos
cos ( )

(23)

which coincides with the static solution in Eqn


(14).
The horizontal component of soil thrust is
determined from the actual geometry, as in the
gravitational problem:

PEH = PE cos( )

111

(24)

predictions are in good agreement (largest


discrepancy about 10%), with the exception of
Coulombs
method
which
significantly
overestimates passive pressures. Moving from
the top to the bottom of each column, an
increase in KA values and a decrease in KP
values can be observed. This is easily
understood given the non-conservative nature
of the first two solutions (Coulomb, Chen), and
the conservative nature of the last two
(Sokolovskii 1965, proposed). This observation
does not hold for the zero extension line
solution of Habibagahi and Ghahramani (1977),
which cannot be classified in the context of limit
analysis theorems. Results for gravitational
active pressures on a rough inclined wall
obtained according to three different methods
as a function of the slope angle , are shown in
Fig 6.

Seismic Component of Soil Thrust

Following Seed & Whitman (1970), the


seismic component of soil thrust is defined from
the difference:

PE = PE P

(25)

which is mathematically valid, as the associated


vectors PE and P are coaxial. Nevertheless, the
physical meaning of PE is limited given that the
stress fields (and the corresponding failure
mechanisms) in the gravitational and seismic
problems are different. In addition, PE cannot
be interpreted in the context of limit analysis
theorems, as the difference of PE and P is
neither an upper nor a lower bound to the true
value.
MODEL VERIFICATION AND RESULTS

Presented in Table 1 are numerical results


for gravitational active and passive pressures
(KA, KP) from the present solution and
established solutions from the literature. The

Table I: Comparison of results for active and passive earth pressures predicted by various methods. The
results for = = 0 are identical for all methods. Note the decrease in KP values as we move from top to
bottom in each column, and the corresponding increase in KA values; = 0 (modified from Chen & Liu, 1990)
a. KA values

20

30

40

20

20

30

30

10

15

20

15

15

0.490

0.447

0.333

0.301

0.217

0.199

0.498

0.476

0.212

0.180

0.490

0.448

0.333

0.303

0.217

0.200

0.498

0.476

0.218

0.189

0.49

0.41

0.33

0.27

0.22

0.17

Slip line (Sokolovskii 1965)

0.490

0.450

0.330

0.300

0.220

0.200

0.521

0.487

0.229

0.206

Proposed Stress Limit


Analysis

0.490

0.451

0.333

0.305

0.217

0.201

0.531

0.485

0.237

0.217

2.04

2.64

3.00

4.98

4.60

11.77

2.27

3.162

5.34

12.91

2.04

2.58

3.00

4.70

4.60

10.07

2.27

3.160

5.09

8.92

2.04

2.55

3.00

4.65

4.60

9.95

Slip line (Sokolovskii 1965)

2.04

2.55

3.00

4.62

4.60

9.69

2.16

3.16

5.06

8.45

Proposed Stress Limit


Analysis

2.04

2.52

3.00

4.44

4.60

8.92

2.13

3.157

4.78

7.07

Coulomb
Kinematic Limit Analysis
(Chen & Liu 1990)
Zero extension (Habibagahi &
Ghahramani 1977)

b. KP values
Coulomb

Kinematic Limit Analysis


(Chen & Liu 1990)
Zero extension (Habibagahi &
Ghahramani 1977)

112

The performance of the proposed solution is


good (maximum deviation from Chens solution
about 10% - despite the high friction angle of
45o) and elucidates the accuracy of the
predictions. The performance of the simplified
solution of Caquot and Kerisel (1948) versus
that of Chen & Liu (1990) is as expected.
Corresponding
predictions
for
passive
pressures are given in Figure 7, for a wall with
negative backfill slope inclination, as a function
of the wall roughness . The agreement of the
various solutions, given the sensitivity of
passive pressure analyses, is very satisfactory.
Of particular interest are the predictions of
Sokolovskiis (1965) and Lee & Heringtons
(1972) methods, which, surprisingly, exceed
those of Chen for rough walls. This trend is
particularly pronounced for horizontal backfill
and values of above approximately 10o and
has been discussed by Chen & Liu.
Results for active seismic earth pressures
are given in Figure 8, referring to cases
examined in the seminal study of Seed &
Whitman (1970), for a reference friction angle of
35o. Naturally, active pressures increase with

0.6
1
K A = PA /( H2 )
2

Coefficient of Active Earth Pressure, K A

0.5

= 45o , = 2 / 3

PA
0.4

= 20o
0.3

Chen & Liu (1990)


Caquot & Kerisel (1948)
Proposed Stress Limit Analysis

0.2

= 0o

0.1

0.0

= 20o

10

15

20

Slope Angle of Backfill,

25
o

Fig 6: Comparison of results for active earth


pressures predicted by different methods
(modified from Chen 1975)

5
1
K P = PP /( H2 )
2

= 30 , = 20

Coefficient of Passive Earth Pressure, K P

PP
4

increasing levels of seismic acceleration and


slope inclination and decrease with increasing
friction angle and wall roughness. The
conservative nature of the proposed analysis
versus the Mononobe-Okabe (M-O) solution is
evident in the graphs. The trend is more
pronounced for high levels of horizontal seismic
coefficient (ah > 0.25), smooth walls, level
backfills, and high friction angles. Conversely,
the trend becomes weaker with steep backfills,
rough walls, and low friction angles.
A similar set of results is shown in Figure 9,
for a reference friction angle of 40o. The
following interesting observations can be made:
First: the predictions of the proposed analysis
are in good agreement with the results from the
kinematic analysis of Chen & Liu (1990), over a
wide range of material and geometric
parameters. Second, the present analysis is
conservative in all cases. Third, close to the
slope stability limit (Fig 9d), or for high
accelerations and large wall inclinations (Fig
9c), Chens predictions are less accurate than
those of the elementary M-O solution.

= 0o

= 10o

= 20o
1
Lee & Herington (1972)
Chen & Liu (1990)
Sokolovskii (1965)
Proposed Stress Limit Analysis

10

20

Angle of Wall Friction,

30

Fig 7: Comparison of results for passive earth


pressures predicted by different methods
(modified from Chen & Liu 1990)

113

interesting comparison is presented in Fig 10b:


average predictions from the two closed-form
solutions (M-O solution & proposed stress
solution) are plotted against the rigorous
numerical results of Chen & Liu (1990).
Evidently, in the range of most practical interest
(30o < < 40o), the discrepancies in the results
have been drastically reduced. This suggests
that the limit equilibrium (kinematic) M-O
solution and the proposed static solution
overestimate and underestimate, respectively,
passive resistances by the same amount in the
specific range of properties. Accordingly, this
averaging might be warranted for design
applications involving passive pressures.
Results for the earth pressure coefficient due
to surcharge KqE (Eq 23) are presented in
Figure 11, for both active and passive
conditions involving seismic action. The
agreement between the stress solution and the
numerical results of Chen & Liu (1990) is
excellent in the whole range of parameters
examined (except perhaps for active pressures,
where ah = 0.3). As expected, M-O solution
performs well for active pressures, but severely
overestimates the passive.

In the same extreme conditions, the


proposed solution becomes exceedingly
conservative, exceeding MO predictions by
about 35%. Note that whereas the M-O and the
proposed solution brake down in the slope
stability limit, Chens solution allows for
spurious mathematical predictions of active
thrust beyond the limit, as evident in Fig 9d.
Fourth,
with
the
exception
of
the
aforementioned extreme cases, Chens and MO predictions remain close over the whole
range
of
parameters
examined.
The
improvement in the predictions of the former
over the latter is marginal.
Results for seismic passive pressures
(resistances) are shown in Figure 10 for the
common case of a rough vertical wall with
horizontal backfill. Comparisons of the
proposed solution with results from the M-O
and Chens kinematic methods are provided on
the left graph (Fig 10a). The predictions of the
stress solutions are, understandably, lower than
those of Chen and Liu, whereas M-O
predictions are very high (i.e., unconservative)
especially for friction angles above 37
degrees. Given the sensitivity of passive
pressure analyses, the performance of the
proposed method is deemed satisfactory. An

114

0.7

0.7

0.6

= 35

= = 0o

0.6

0.5

= 35o

0.5

= 0o

= 0o

=/2

0.4

K E cos

Coefficient of Seismic Active Earth Pressure, K E

= = 0o

0.3

0.2

0.4

=/2
0.3

0.2

0.1

0.1
M - O Analysis
Proposed Stress Limit Analysis

0.0
0.0

0.1

0.2

0.3

0.4

M - O Analysis
Proposed Stress Limit Analysis

0.0
0.0

0.5

0.1

0.3

0.4

0.5

Horizontal Seismic Coefficient, ah

Horizontal Seismic Coefficient, ah

0.7

0.7

= = 0o

= = 0o
0.6

0.2

= 30

=/2

0.6

= 35 o ; = / 2
o

= 20

3 5o

4 0o

0.5

0.5

K E cos

K E cos

=0

0.4

0.3

0.2

0.4

0.3

0.2

0.1

0.1
M - O Analysis

M - O Analysis
Proposed Stress Limit Analysis

Proposed Stress Limit Analysis

0.0
0.0

0.1

0.2

0.3

0.4

0.5

Horizontal Seismic Coefficient, ah

0.0
0.0

0.1

0.2

0.3

0.4

Horizontal Seismic Coefficient, ah

Fig 8: Comparison of active seismic earth pressures predicted by the proposed


solution and from conventional M - O analysis, for different geometries, material
properties and acceleration levels ; av = 0. (modified from Seed & Whitman, 1970)

115

0.5

1
K AE = PAE /( H2 )
2
= = 0o ; = 2 / 3

(a)

0,6

(b)

PAE

0,5

0,5

0.30
0.20

0,4

30

35

40

M - O Analysis
Kinematic Limit Analysis (Chen & Liu 1990)
Proposed Stress Limit Analysis

0,1

45

0,0

0,1

0,2

0,3

0,4

Horizontal Seismic Coefficient, ah

Friction Angle, o

1,6

1,0

K AE

(d)

1
= PAE /( H2 )
2

Coefficient of Seismic Active Earth Pressure, K AE

(c)

= 40 ; ah = 0.20 ; = / 2
0,8

PAE
15o

0,4

= 0o
15o

0,2
M - O Analysis
Kinematic Limit Analysis (Chen & Liu 1990)
Proposed Stress Limit Analysis

-20

/ =1

0,2

0,1
25

Coefficient of Seismic Active Earth Pressure, K AE

M - O Analysis
Kinematic Limit Analysis (Chen & Liu 1990)
Proposed Stress Limit Analysis

0,0

0,4

ah = 0

0,2

0,6

PAE

0,3

0.10
0,3

1
K AE = PAE /( H2 )
2
= = 0o ; = 40o

0,6

K AE cos

Coefficient of Seismic Active Earth Pressure, K AE

0,7

-10

10

Slope Angle of Backfill,

1,4

1,0

slope
stability

= 40o ; = 0o ; = / 2

1,2

limit

PAE

0,8

=/2

0,6

/3
0o

0,4

0,2

0,0
0,0

20

1
K AE = PAE /( H2 )
2

M - O Analysis
Kinematic Limit Analysis (Chen & Liu 1990)
Proposed Stress Limit Analysis

0,1

0,2

0,3

Horizontal Seismic Coefficient, ah

Fig 9: Comparison of active seismic earth pressures predicted by different methods,


for different geometries, material properties, and acceleration levels; f = 40o,
av = 0. (modified from Chen & Liu, 1990)

116

0,4

25

25

Coefficient of Seismic Passive Earth Pressure, K PE

(a)

20

15

(b)

H PPE

H PPE

20

ah = 0
0.1

1
K PE = PPE /( H2 )
2

0.2
0.3

= 0o , = 0o
= 2 / 3

10

10

30

35

ah = 0
0.1
0.2
0.3

Kinematic Limit Analysis (Chen & Liu 1990)


Proposed Stress Limit Analysis

0
25

1
K PE = PPE /( H2 )
2

15

Mononobe - Okabe
(ah = 0)

Kinematic Limit Analysis (Chen & Liu 1990)


Average of M - O & Proposed Stress Limit Analysis

40

0
25

45

30

35

40

45

Angle of Internal Friction, o

Angle of Internal Friction, o

Fig 10: Comparison of results for passive seismic resistance on a rough wall
predicted by various methods. (Modified from Chen & Liu, 1990)

20

0,6

= = 0o

PAE

0,5

ah q
H

0.3
0.2

0,4

0.1
0,3

PPE

15

=2/3

K AE = PAE / q H

(b)

ah q

K PE = PPE / q H
q

0,7

(a)

10

ah = 0
0.1
0.2
0.3

= = 0o
=2/3

Mononobe - Okabe
(ah = 0)

ah = 0
5

0,2
Kinematic Limit Analysis (Chen & Liu 1990)
Proposed Stress Limit Analysis

0,1
25

30

35

Friction Angle,

40

Kinematic Limit Analysis (Chen & Liu 1990)


Proposed Stress Limit Analysis

45

0
25

30

35

Friction Angle,

40
o

Fig 11: Variation of KAEq and KPEq values with - angle for different acceleration levels.

117

45

Distribution of Earth Pressures on the Wall:


Analytical Findings

cos = 1 + tan 2 stands for the vectorial sum


of shear and normal tractions at the wall-soil
interface. Factor cos arises from the
equilibrium of the infinite slope in Eqn (4a).
Finally, cos( ) / cos2 is a geometric factor
arising from the integration of stresses along
the back of the wall, and is associated with the
inclination of the wall and backfill.
In light of
the above, the solution for gravitational
pressures can be expressed by the single
equation

Mention has already been made that in the


realm of pseudo-dynamic analysis, there is no
fundamental physical difference between
gravitational and seismic earth pressures.
Equations (4) indicate that stresses in the soil
vary linearly with depth (stress fan does not
alter this dependence), which implies that both
gravitational and seismic earth pressures vary
linearly along the back of wall. In the absence
of surcharge, the distribution becomes
proportional with depth, as in the Rankine
solution. Accordingly, the point of application of
seismic thrust is located at a height of H/3
above the base of the wall. It is well known from
experimental
observations
and
rigorous
numerical solutions, that this is not generally
true. The source of the difference lies in the
distribution of inertial forces in the soil mass
(which is often sinusoidal like - following the
time-varying natural mode shapes of the
deposit), as well as the various kinematic
boundary conditions (wall flexibility, foundation
compliance, presence of supports). Studying
the above factors lies beyond the scope of this
article, and like will be the subject of a future
publication. Some recent developments are
provided in the Master thesis of the second
author (Kloukinas, 2006) as well as in (Kramer,
1996;
Veletsos
&
Younan,
1994;
Theodorakopoulos et.al. 2005; Ostadan, 2001;
Paik & Salgado, 2003) .

K =

cos ( ) cos
cos cos2

1 sin cos ( 2 )

exp( 2 tan )
1 + sin cos[ 1 + ]

(26)

which is valid both for active conditions (using


positive values for and ) and passive
conditions (using negative values for and ). It
is straightforward to show that this property is
not valid for the Mononobe-Okabe solutions in
Eqn (4). The lack of symmetry in the limit
equilibrium solutions can be attributed to the
optimization
(i.e.,
maximization
and
minimization) involved in deriving the limit
thrusts. An application example elucidating the
simplicity of the solution is provided below.
Application Example

Active and passive pressures will be


computed for a gravity wall of height H = 5m,
inclination = 5o and roughness = 20o,
retaining an inclined cohesionless material with
= 30o, =18kN/m3 and = 15o, subjected to
earthquake accelerations ah = 0.2 and av = 0.
The static counterpart of the problem has been
presented by Powrie (1997).
The inclination of the resultant body force in
the backfill is obtained from Eqn (1):

DISCUSSION: SIMPLICITY & SYMMETRY

It is instructive to show that the proposed


solution can be derived essentially by
inspection. Indeed, basis of Equation (12) is the
familiar Rankine ratio (1 sin) / (1 sin). The
terms cos(2 ) and cos (1 ) in the
numerator and denominator reflect the fact that
stresses and w are not principal. Both terms
involve the same double signs as their
multipliers (i.e. sin and sin, respectively).
Angle and associated Caquot angle 1 have
to be in the denominator, as an increase in their
value must lead to an increase of active thrust.
The exponential term is easy to remember and
involves the same double sing ( ) as the
other terms in the numerator. With reference to
the
factors
outside
the
brackets,

e = Arctan(0.2) = 11.3o

(A-1)

The two Caquot angles are determined from


Eqns (6), (9) and (15) as

1* = sin 1[sin (15 + 11.3 ) / sin 30] = 62.4o

118

(A-2)

2 = sin [sin ( 20 ) / sin 30] = 43.2


1

PAH = 94.5 cos(5 + 20) = 85.6 kN / m


o

(A-3)
Note that according to Powrie (1997), the
value of the horizontal component is (Eqn 9.42,
p 333):

The angle separating the major principal


planes in Regions A and B is computed from
Eqn (22):

2 E = 43.2 ( 62.4 + 20 ) + 15 2 5 11.3


= 45.5o

1
0.395 18 52 (1 + tan 5 tan 20 )
2
= 91.7 kN / m

PAH =
(A-4)

which is in error as: (1) Ka, as determined from


Eqn 9.41 in Powries book, should be 0.385
not 0.395. (2) the correct sign in front of product
(tan x tan) is minus one. (3) Powries
equation does not encompass factor [cos( )
/ cos cos] arising from the integration of
stresses on the back of the wall.
For the passive case, the corresponding
parameters are:

Based on the above values, the earth


pressure coefficient is obtained from Eqn (21):

K AE =

cos ( 5 15) cos(15 + 11.3)


cos11.3cos20 cos2 5

1 sin 30 cos ( 43.2 20 )

1 + sin 30 cos 62.4 + (15 + 11.3)


exp(+ 45.5

180

(A-5)

e = Arctan(0.2) = 11.3o ,

tan 30) = 0.82

1* = sin 1[sin (15 11.3) / sin 30] = 7.4o ,


2 E = 43.2 + ( 7.4 + 20 ) + 15 2 5 + 11.3 = 86.9o ,

from which the overall active thrust on the wall


is easily determined (Eqn 11):

PAE =

1
0.82 18 52 = 185 kN / m
2

For
the
gravitational
corresponding parameters are

problem,

(A-8)

The passive earth pressure coefficient and


resistance are obtained from Eqns (21) and
(11):

(A-6)

K PE =

cos ( 5 15) cos(15 11.3)


cos11.3cos20 cos2 5

1 + sin 30 cos ( 43.2 + 20 )

1 sin 30 cos 7.41 (15 11.3)

the

1 = sin1[sin15 / sin 30] = 31.2o ,

exp(+ 2 E

2 = sin 1[sin ( 20 ) / sin 30] = 43.2o ,

180

(A-9)

tan 30) = 6.31

2 = 43.2 ( 31.2 + 20 ) + 15 2 5 = 3o ,

K A = 0.42 .

PPE =

Thus,

PA =

1
0.42 18 52 = 94.5 kN / m
2

1
6.3118 52 = 1420 kN / m
2

(A-10)

CONCLUSIONS

(A-7)

A stress plasticity solution was presented for


determining gravitational and earthquakeinduced earth pressures on gravity walls
retaining cohesionless soil. The proposed
solution incorporates idealized, yet realistic wall

The horizontal component of gravitational


soil thrust is determined from Eqn (24)

119

geometries and material properties. The


following are the main conclusions of the study:

appropriate signs for friction angle and wall


roughness.

1) The proposed solution is simpler than the


classical Coulomb and Mononobe-Okabe
equations. The main features of the
mathematical expressions, including signs, can
be deduced by physical reasoning, which is
hardly the case with the classical equations.

It should be emphasized that the verification


of the proposed solution was restricted to
analytical and not experimental results. Detailed
comparisons against experimental results
including distribution of earth pressures along
the wall, will be the subject of a future
publication.

2) Extensive comparisons with established


numerical solutions indicate that the proposed
solution is safe, as it overestimates active
pressures and under-predicts passive. This
makes the method appealing for practical
applications.

ACKNOWLEDGMENTS

The Authors are indebted to Professor


Dimitrios Atmatzidis for his constructive
criticism of the work.

3) For active pressures, the accuracy of the


solution is excellent (maximum observed
deviation from numerical data about 10%). The
largest deviations occur for high seismic
accelerations, high friction angles, steep
backfills, and negative wall inclinations.

REFERENCES
Anastasopoulos I., Gazetas G., Psarropoulos Pr.
(2003), Flexible Retaining Walls : Why They do
not often Fail in Strong Seismic Shaking, Proc.
Fib Int. Symposium on Concrete Structures in
Seismic Regions, Athens, May 2003.

4) For passive resistances, the predictions are


also satisfactory. However, the error is larger especially at high friction angles. Nevertheless,
the improvement over the M-O predictions is
dramatic. Taking the average between the
predictions of the M-O solution and the
proposed stress solution yields results which
are comparable to those obtained from rigorous
numerical solutions.

Atkinson, J. (1981), Foundations and slopes,


McGraw Hill, London.

5) The pseudo-dynamic seismic problem can


be deduced from the corresponding static
problem through a revolution of the reference
axes by the seismic angle e (Fig 5). This
similarity suggests that Coulomb and M-O
solutions are essentially equivalent.

Chen, W.F. (1975), Limit analysis and soil plasticity,


Developments in geotechnical engineering,
Elsevier, Amsterdam.

Briske, R. (1927), "Die Erdbebensicherheit von


Bauwerken", Die Bautechnik, Vol 5, 425- 430,
453-457, 547-555.
Caquot, A. and Kerisel, L. (1948), Trait de
mcanique des sols, Gauthier-Villars, Paris.

Chen, W.F., Liu, X.L. (1990), Limit analysis in soil


mechanics, Elsevier, Amsterdam.

6) Contrary to the overall gravitational-seismic


thrust PE, the purely seismic component PE =
PE P cannot be put in the context of a lower or
an upper bound. This holds even when PE and
P are upper or lower bounds.

Clough GW, Duncan JM. (1990), Earth pressures. In


Foundation Engineering Handbook, Fang HY
(ed.). Chapman & Hall:New York, pp.223-235.
Collins IL. (1973), A note on the interpretation of
Coulomb analysis of the thrust on a rough
retaining wall in terms of the limit theorems of
plasticity theory, Geotechnique, Vol. 24, No 1,
pp.106-108.

7) In the realm of the proposed model, the


distribution of earth pressures on the back of
the wall is linear with depth for both
gravitational and seismic conditions. This is not
coincidental given the similarity between the
gravitational and pseudo-dynamic problem.

Coulomb, C.A. (1776), "Essai sur une application


des regles de maximis et minimis a quelqes
problemes de stratique relatifs a l architecture".
Memoires de Mathematique et de Physique.
Presentes a l Academie Royale des Sciences;
Paris, 7, 343-382.

8) Contrary to the Mononobe-Okabe solutions,


the proposed solution is symmetric, as it can be
expressed by a single equation, describing both
active and passive pressures, by means of

120

Davis, R.O. and Selvadurai, A.P.S. (2002), Plasticity


and Geomechanics, Cambridge University Press,
Cambridge.

Lambe TW, Whitman RV. (1969), Soil Mechanics,


Wiley & Sons: NY.
Lancelotta R. (2002), Analytical solution of passive
earth pressure. Geotechnique, 52, 8, pp. 617619.

Davies TG, Richards R, Chen KH (1986). Passive


pressure during seismic loading, Journal of
Geotechnical Engineering, Vol. 112, No 4, pp.
479-483

Lee, I.K. and Herington, J.R. (1972), "A theoretical


study of the pressures acting on a rigid wall by a
sloping earth or rock fill", Geotechnique, 22, No
1, 1-26.

Ebeling, R.M., Morrison, E.E., Whitman, R.V., Liam


Finn, W.D. (1992), A Manual for Seismic Design
of Waterfront Retaining Strutures, US Army
Corps of Engineers, Tech. Report ITL-92-11.

Matsuo M, Ohara S. (1960), Lateral earth pressures


and stability of quay walls during earthquakes,
Proceedings, Second World Conference on
Earthquake Engineering, Tokyo, Japan.

Finn WD. Applications of Limit Plasticity in Soil


Mechanics. (1967), Journal of the Soil Mechanics
and Foundations Division, ASCE, Vol. 93 (SM5),
pp.101-120.

Mononobe, N., Matsuo, O. (1929), On the


determination
of
earth
pressure
during
earthquakes,
Proceeding
of
the
World
Engineering Congress, Tokyo, vol 9, 179-187.

Finn WD, Yogendrakumar M, Otsu H, Steedman


RS. (1989), Seismic response of a cantilever
retaining wall: centrifuge model test and dynamic
analysis. Proceedings of 4th International
Conference on Soil Dynamics and Earthquake
Engineering.
Computational
Mechanics
Publications: Southampton, pp.331-431.

Okabe, S. (1924), "General theory on earth pressure


and seismic stability of retaining walls and dams",
Journal of the Japanese Society of Civil
Engineers, 10, 6, 1277-1323.
Ostadan F. (2005), Seismic soil pressure for building
walls: an updated approach, Soil Dynamics and
Earthquake Engineering, 25, pp 785-793

Garini, E., Gazetas, G., and Gerolymos, N. (2006),


Asymmetric block sliding from idealized pulse
wavelets and near-fault ground motions,
Panhellenic Conference on Geotechnical &
Geoenvironmental Engineering, Xanthi, May 31
June 2 (in Greek).

Paik K and Salgado R. (2003), Estimation of Active


Earth Pressure Against Rigid Retaining Walls
Considering Arching Effects, Geote-chnique, 53,
No 7, pp 643-653.

Gazetas G., Psaropoulos P.., Anastasopoulos I.,


and Gerolymos N. (2004), Seismic Behaviour of
Flexible Retaining Systems Subjected to Short
Duration Moderately Strong Excitation, Soil
Dynamics & Earthquake Engineering, Vol. 24,
No. 7, pp. 537-550.

Papantonopoulos, C. and Ladanyi, B. (1973),


"Analyse de la Stabilitee des Talus Rocheux par
une Methode Generalisee de lEquilibre Limite",
Proceedings, 9th Canadian Rock Mechanics
Symposium, Montreal, 167-196 (in French).

Gazetas G., Gerolymos N., Anastasopoulos I.


(2005), Response of Three Athens Metro
Underground Structures in the 1999 Parnitha
Earthquake, Soil Dynamics & Earthquake
Engineering, Vol. 25, No. 7-10, pp. 617633.

Parry RHG. (1995), Mohr Circles, Stress Paths and


Geotechnics, E & FN Spon.
Pecker A. (1995), Seismic design of shallow
foundations. State-of-the-Art: 10th European
Conference on Earthquake Engineering, Duma
Ed. Balkema, pp 1001-1010.

Habibagahi K and Ghahramani A. (1977), Zero


extension theory of earth pressure. Journal of the
Geotechnical Engineering Division, ASCE;
105(GT7): pp.881-896.

Powrie W. (1997), Soil Mechanics: Concepts &


Applications, E & FN Spon, London.
Psarropoulos PN, Klonaris G, Gazetas G. (2005)
Seismic earth pressures on rigid and flexible
retaining walls, Soil Dynamics and Earthquake
Engineering; Vol. 24, p.p. 795809.

Heyman J. (1972), Coulombs Memoir on Statics; an


essay in the history of civil engineering,
Cambridge University Press
Kramer, S.L., (1996), Geotechnical Earthquake
Engineering, Prentice Hal.

Richards R, Elms DG. (1979), Seismic behaviour of


gravity retaining walls. Journal of the
Geotechnical Engineering Division, ASCE;
105(GT4): pp.449-464.

Kloukinas, P., (2006), "Gravitational and Seismic


Earth Pressures by Stress Limit Analysis", Ms
Thesis, University of Patras, Greece (in Greek
with extended English summary).

Salencon, J. (1974), Applications of the Theory of


Plasticity in Soil Mechanics. John Wiley.

121

Salencon, J. (1990), Introduction to the yield design


theory, European Journal of Mechanics A/Solids,
Vol. 9, No 5, pp. 477-500

Theodorakopoulos DD, Chassiakos AP, Beskos DE.


(2001), Dynamic pressures on rigid cantilever
walls retaining poroelastic soil media. Part II:
Second method of solution, Soil Dynamics &
Earthquake Engineering, 21, 4, p. 339-364.

Seed HB, Whitman, R.V. (1970), Design of earth


retaining
structures
for
dynamic
loads,
Proceedings of specialty conference on lateral
stresses in the ground and design of earth
retaining structures, ASCE, Ithaca, New York,
103-147.

Veletsos A.S. and Younan A.H. (1994). "Dynamic


soil pressures on rigid retaining walls",
Earthquake
Engineering
and
Structural
Dynamics, 23, pp. 275 301.

Sheriff MA, Ishibashi I, Lee CD. (1982), Earth


pressures against rigid retaining walls, Journal
of Geotechnical Engineering, ASCE ; 108: GT5,
pp. 679-696.

Whitman, R.V. and Liao, S. (1985), Seismic Design


of Gravity Retaining Walls, US Army Corps of
Engineers, Miscellaneous paper GL-85-1.
Wood, JH. (1973), Earthquake induced soil
pressures on structures. Doctoral Dissertation,
EERL 73-50, California Institute of Technology,
Pasadena, CA.

Sokolovskii, V.V. (1965), Statics of granular media,


Pergamon Press, New York.
Steedman RS, Zeng X. (1990), The influence of
phase on the calculation of pseudo-static earth
pressure on a retaining wall, Geotechnique, Vol.
40, pp.103-112
Terzaghi, K., (1943), Theoretical soil mechanics,
John Wiley & Sons Inc., New York.
Theodorakopoulos DD, Chassiakos AP, Beskos DE.
(2001), Dynamic pressures on rigid cantilever
walls retaining poroelastic soil media. Part I: First
method of solution, Soil Dynamics & Earthquake
Engineering, 21, 4, p. 315-338.

122

A Study on Ground Displacement due to Fault Slip


and its Influence to Underground Structure
Y. Adachi1, S. Yoshimura , T. Nakata
1
2

Hanshin Expressway Company Limited, Kyoto, Japan


Yachiyo Engineering Company Limited, Niigata, Japan

Abstract
The influence to shield tunnel structure due to ground slip, crack, and deformation
propagation in sedimentary layer caused by fault slip is studied. The base rock
displacement is estimated by empirical relations between the earthquake magnitude and the
base rock displacement. The ground displacement propagation is analyzed using linear and
non-linear finite element method with joint elements. As the result of this study, it is found
that the fault displacement does not reach to the surface where the base rock displacement
is relatively smaller than the thickness of the sedimentary layer. Therefore, in the case
where the tunnel existed in relatively shallow depth in deep sedimentary layer, critical
damage may not occur to the tunnel even though the tunnel suffers the ground
displacement.

between the experiment results and the site


survey results. Ramancharla and Megro (2001)
estimated the propagation of fault rupture in
deposit layer due to base layer slip by applied
element analysis. Konagai et al. (2001) also
simulated the propagation in deposit layer by
Lagrangian particle finite difference method.
There are some studies about the fault
rupture propagation in soil and its influence to
structures. Newmark et al (1975) proposed the
design method of buried pipeline to resist
against large fault displacement. Kennedy et al
(1977) estimated fault movement effects on
buried oil pipeline. Kiyomiya et al (2003)
calculated the fault movement effects on buried
sea bed pipeline by beam elements on soil
spring. Takada et al (2001) studied the relation
between the strain on buried pipeline and fault
rupture propagation and proposed the
estimation method of the maximum strain of the
pipeline. As described above, almost of the
researches were based on the ideal condition,
not the realistic condition.
The authors studied the shield tunnel
designed in the diluvium soil layer with crashed
planes
caused
by
fault
displacement
propagation by finite element analysis using

INTRODUCTION

A lot of social infrastructure such as bridges


and tunnels were suffered by the ground
displacement due to the fault slip of 1999
Kocaeli earthquake and Chi-Chi earthquake. It
is generally said very difficult to design resistant
structures against such ground displacement,
especially underground structures. However it is
very important to estimate the damage due to
such ground displacement and take possible
fail safe measures as much as possible.
There are a lot of studies about the fault
displacement propagation in deposit layer. Cole
et al. (1984) discussed influence zones in
alluvium over dip-slip faults by sand box
experiment. Bray et al. (1994) studied
earthquake fault rupture propagation through
soil due to reverse fault displacement by
nonlinear FEM analysis and sand box
experiment. Taniyama et al. (1998) estimated
the deformation in sandy deposit layer due to
reverse faulting by nonlinear FEM analysis and
sand box experiment. Onizuka et al. (1999)
studied the propagation of reverse faulting in
the sandy deposit layer by aluminum bar
experiment and showed the good agreement

123

600

550

500

450

400

350

300

100

250

200

150

100

50

-150

-200

-250

-300

-350

-400

METER

METER

METER

Crashed Planes
Shield Tunnel

100

Surface

Alluvium Layer

Diluvium Layer
-100

Base Line

-100

Base Rock

-200

-200

-300

-300

-400

-400

Fig. 1: Ground condition, location of crashed planes, and location of shield tunnel structure
Table 1: Estimated ground displacement

joint elements. This paper studies the difference


of the fault displacement propagation on linear
or non-linear soil properties and also discusses
the difference of the influence to the tunnel
structure.

Source

EQ Type

Ground

Matsuda

Inland (Japan)

Surface

Satoh et al

All (Japan)

Subface

69.2

Ave

Takemura et al

Inland (Japan)

56.5

Ave

SOILCONDITION AND SHIELD TUNNEL


Donald et al

The soil condition and the shield tunnel


structure studied in this paper are shown in Fig
1. Reflection test was performed in this field
and four crashed planes were found in the
deposit sediment layer indicated by ,,,
and in Fig. 1. There also found the fault
flexure lines at the ground surface related to
those crashed planes. The faults were
considered as reverse faults, and the return
period was estimated around 10,000 years, and
the maximum possible magnitude was
estimated M=6.5 by the area of the fault plane.
The planned tunnel was a shield tunnel with
composite segments. The diameter of the
tunnel was 10.6m. The aliment of the tunnel
was not so deep in the ground because of the
road tunnel.

Displ.(cm)
79.4

Ave

All (World)

Surface

55.4

Max

All (World)*

Surface

90.1

Max

All (World)

Surface

29.9

Ave

All (World)*

Surface

8.4

Ave

*: Only reverse faults are considered

displacement at the base layer.


Current studies use the statistical estimation
method for determining the base layer
displacement by analyzing the relation between
the displacement of the surface ground and the
magnitude of the earthquakes. These statistical
studies do not focused on the presence of the
deposit layer so that the obtained statistical
relation is not direct relation between the base
layer displacement and the earthquake
magnitude. But in this study, the statistical
estimation method can be assumed to estimate
the base layer displacement. Freeman et al
(2000) also took the same assumption and
studied about the assessment of the dam
against the fault displacement.
In this study, the works of Matsuda (1975)
which studied the fault displacement by
collecting active faults earthquake data
obtained in Japan, that of Sato et al (1989)
which studied about the active and plate
boundary earthquake in and around Japan, that
of Takemura et al (1998) which studied about
the in-crust earthquake, and that of Donald et al
(1994) which studied about the all of the

ESTIMATION OF FAULT DISPLACEMENT AT


BASE LAYER

First of all, the fault displacement at the


base layer is needed to be determined in order
to study the fault displacement propagation in
the deposit layer. Generally said, the base layer
displacement and the fault displacement are
different. Therefore the fault displacement used
for the strong motion analysis by fault rupture
modeling is not good for estimating the

124

Table 2: Soil properties of deposit layers


3

+40m

(kN/m ) VS(m/s) VP(m/s)


Alluvium
15.7
200
1,600
16.7
400
1,700
Diluvium

Crashed Plane()

18.6

Diluvium

Crashed Plane()

600

G(kN/m ) E(kN/m )
0.49 63,800 192,000
0.47 273,000 800,000
1,980,00
0.44 695,000
0

1,900

Shield Tunnel

Table 3: Joint element


Deposit Layers

m
Alluviu

Layer

m
Diluviu

Layer

Diluvium

as
Cr

es
lan
dP
he

C(kN/m )
0
0

Model 1
Model 2

Base Displacement

Layer

KnkN/m
1,000
1,000

Table 4: Physical properties of tunnel segment

yer
Base La

(kN/m )

Im

EkN/m

24.0

0.15

108.9

486,000

Equivalent
stiffness

Fig. 2: Modeling of ground and tunnel for FEM analysis

earthquake recorded in the world so far are


used
for
estimating
the
base
layer
displacement.
Table 1 shows the result of the base layer
displacement obtained by those studies above.
The target fault is considered as inland active
fault so that the study by the Matsuda (1975) is
seemed good to be applied in this study. By
using the Matsuda relation, 80cm of the base
layer displacement is computed for the target
fault. However, the Matsuda formula computes
the average value, so that the dispersion of
fault rupture phenomenon should be considered
from the engineering point of view. In this study,
10% of the dispersion is considered. So finally,
90cm of base layer displacement is used for
computation.
Fig 1 shows the four found crashed planes.
From the fault geology point of view, the most
distant crashed plane in the axis of fault incline
direction is the newest. But the cover depth of
surface layer is the deepest in the four planes.
Therefore, the oldest crashed plane where the
cover depth is the shallowest among the four is
employed in this study.

KrkN/m
1,000
1,000

35
20

Table 5: Study cases


Soil model
Linear
Linear
Non-linear
Non-linear

Case 1
Case 2
Case3
Case4

Crashed plane
Joint model 1
Joint model 1
Joint model 1
Joint model 1

n Normal stress

Tension

Tunnel stiffness
Equivalent Stiff.
Equivalent Stiff.
Equivalent Stiff.
Equivalent Stiff.

Shear stress
y=|n|tan+c

Sliding

Ks

Kn
Compression

Shear strain

Normal strain
Open

Sliding
-y=-|n|tan+c

(1) Normal direction


(2) Sliding direction
Fig. 3: Modeling of Crashed plane

is modeled by 3 layers, one alluvium, and two


diluvium layers. The perpendicular section in
which the shield tunnel is existed is used for the
two dimensional analysis.
As for the soil modeling, linear solid
elements are used for the three-dimensional
analysis and nonlinear solid model which
assumed the bilinear behavior that showed the
Mohr-Coulomb yield relation. The density , P
wave velocity Vp, and S wave velocity Vs are
computed from the result of geophysical logging.
The dynamic Poisson ratio , shear modulus G,
elastic modulus E are computed the estimated value
above mentioned.

ANALYSIS METHOD AND MODELING

Joint element which can express slide and


opening is employed for expressing the
geophysical property of the crashed plane. The
modeling of the joint element was shown in Fig.
3. Table 3 shows also the modeling of the joint
element. The cohesion is assumed c=0 for the
safety side computation. The internal friction
angle is assumed =35which is the standard
value for sand layer, and for=20which is

In order to study the fault rupture


propagation in deposit layer, three-dimensional
FEM analysis are performed with linear or nonlinear solid elements, joint elements with
standard value or with limited value.
Fig. 2 shows the FEM model used for the
three dimensional analysis. The analyzed area
is 720m long and 200m wide. The deposit layer

125

the reduced value which considers the repeated


fault slide.
For the shield tunnel, the tunnel is modeled
as a beam element which has effective stiffness
considering the ring joint stiffness.
As for the boundary condition, the base
layer is fixed. 90 cm of base layer displacement
with 45 incline angle is applied at the upper
(right) side of the base layer as shown in Fig. 1
and 2.
Table 5 shows the analytical cases
conducted in this study. All the cases, joint
elements are used for the crashed planes.
Case 1 is the case of three-dimensional
analysis with linear solid model. Case2 is the
two dimensional analysis with linear solid model.
Case 3 is the two dimensional analysis with
nonlinear solid model. Case 4 is the same as
case 3 except the internal friction angle
=20for the joint elements.
RESULT OF THE ANALYSES

Fig. 4 shows all the results of the


displacement of soil after being applied 90cm
offset at the base layer. Case1 shows the result
of the tunnel plane for the comparison. The
vertical direction is amplified 20 times of actual
scale. No evident gaps are observed along the
crashed plane for all the results even though all
the cases use joint elements for them.
Focused on the surface gap at the ground
surface indicated by the yarrows in Fig. 4,
2.1cm for Case 1, 2.1cm for Case 2, 0.0cm for
Case 3, and 0.0 cm for Case 4 are computed
so that the slide effects are so small compared
to the 90cm of the base offset. It can be
concluded that almost all the base offset are
absorbed in the soil in these cases. This agrees
with the condition that no evident fault line, only
flexure line, is observed at the field surface.
According to the previous studies, fault rupture
propagation can reach to the ground where the
base offset is over 4 to 5% of the cover
thickness of the deposit layer (Bray et al (1992)).
This case is the case where only 1% offset of
the deposit layer is applied to the deposit layer
so that the result obtained by this study agrees
with that of the previous ones.
Focused on the deformation mode of the
deposit layer, no difference to the Case1 and 2
was observed so that it can be said that there is
no clear difference to three and two

126

dimensional analyses. As for Case 1, the clear


difference is not observed for the deformation in
the tunnel plane and the side boundary plane
so that it can be said the tunnel stiffness is too
small compared to the soil.
There are clear difference of the soil
deformation modes compared to the results of
Case2, 3 and 4. As for Case2 where linear solid
model is used, only the upper side of the
deposit layer is displaced following with the
base offset. As for Case 3 and 4, not only upper
side but also lower side of the deposit layer
deform.
Fig. 5 shows the horizontal and vertical
displacement contour map of Case 2 and 3.
According to the result of Case 2, the lower side
of the deposit layer does not move in the
vertical direction but move in the horizontal
direction. The influenced area of Case 3 is
developed compared to that of Case 2. The
influence is concentrated to the joint element
where the linear solid model is used, but the
influenced area is widely developed where the
nonlinear solid model is used.
Fig. 6 shows the status of the joint elements,
Open, Slide, and Close. As for Case 2,
almost all elements of the crashed plane
show Slide except showing Open at the
bottom. The rest of the crashed plane show
Close so that it can be seen that the
displacement
is concentrated to just one
location. As for Case3 where the soil is
modeled by nonlinear elements so that the
lower deposit layer also deforms, the status of
the crashed planes are similar to those of Case
2. As for Case 4 where internal friction angle is
reduced, the crashed plane where the base
offset is applied and the adjacent plane show
Sliding.
Fig. 7 shows the relative displacement
between the upper and lower deposit layer
along the crashed plane of all the cases. As
for Case1 and 2, the influence of the offset of
the base layer reaches to about 30m upward
from the base which is almost 30 times of the
amount of the offset. The difference between
the two cases is pretty small. As for Case3 and
4 where the soil condition is modeled by
nonlinear elements, almost all the base offset is
absorbed with in 10m above the base. In these
cases, the base offset is absorbed in the lower
part of the deposit layer, not concentrated to the
joint elements.

(a) Case1

(b) Case2

(c) Case3

(d) Case4

Fig 4: Ground deformation (Red yarrows indicate the locations at the ground surface gaps)
0.

0.0

0.

0.0

-0.1

0.1

-0.1

0.1

-0.2

0.2

-0.2

0.2

-0.3

0.3

-0.3

0.3

-0.4

0.4

-0.4

0.4

-0.5

0.5

-0.5

0.5

-0.6

0.6

-0.6

0.6

-0.7

0.7

-0.7

0.7

-0.8

0.8

-0.8

0.8

-0.9

0.9

-0.9

0.9

-1.

1.0

Output Set: Interval 1, Step


Contour: X Translation

-1.

(a) Horizontal displacement contour map (Case2)

1.0Output Set: Interval

1, Step 100
Contour: X Translation

(b) Horizontal displacement contour map (Case3)Unit(m)

Unit(m)
1.

1.0

1.

1.0

0.9

0.9

0.9

0.9

0.8

0.8

0.8

0.8

0.7

0.7

0.7

0.7

0.6

0.6

0.6

0.6

0.5

0.5

0.5

0.5

0.4

0.4

0.4

0.4

0.3

0.3

0.3

0.3

0.2

0.2

0.2

0.2

0.1

0.1

0.1

0.1

0.

0.0

Output Set: Interval 1, Step


Contour: Y Translation

(c) Vertical displacement contour map (Case2)

Output Set: Interval 1, Step 100


Contour: Y Translation

Unit(m)

(d) Vertical displacement contour map (Case3))

Fig. 5: Horizontal and vertical displacement contour map

127

0.

0.0
Unit(m)

0
Case1
Case2
Case3
Case4

Depth from the Ground(m)

-10

Open 3.
Sliding2.
Close

Y
Z

1.

0.

-20

Tunnel location

-30
-40
-50
-60
-70
-80

(a) Case2

-90
0

20

40
60
80
Relative Displacement(cm)

100

Fig. 7: Relative displacement along the crashed


plane where the base offset is applied
0.7
0.6
3.

Vertical displacement (m)

Open 2.
Sliding
Close 1.

0.

(b) Case3

Location of
Crashed Planes

0.5
0.4
0.3

Case3

Case2

0.2
0.1
0
-0.1

200

400

600

Distance(m)
3.

Y
Z

Fig. 8: Vertical displacement along tunnel axis

Open 2.
Sliding1.
Close

consideration is needed for studying the cases


using nonlinear soil condition.
Fig. 8 shows the longitudinal displacement
of the tunnel of Case2 and 3. The lateral axis
shows the distance from the left edge of the
model. As for Case 2, the big displacement
jump occurs at the crashed plane where the
base offset is applied. But on the contrary, as
for Case 3, the displacement jump occurs at the
left area from the offset crashed plane.

0.

(c) case4
Fig. 6: Status of joint elements for crashed planes

The previous study by Bray et al (1992), shows


the fault rupture propagation can reach the
ground surface when the amount of the base
offset is over 3 to 6% of the depth of the deposit
layer. These research results show good
agreement to the result of Case 1 and 2 and not
to the result of Case 3 and 4. This is because
the previous studies assume the linear soil
property, but the Case3 and 4 assume
nonlinear soil behavior. Therefore more

INFLUENCE TO TUNNEL STRUCTURE

Using the ground displacement analyzed


above, the influence to the tunnel structure due
to the ground displacement caused by fault

128

maximum moment is observed at the left side of


the offset plane. These observations and
tendencies are similar to the vertical
displacement of the tunnel.
In any case. The relative displacement
along the offset plane is absorbed below the
tunnel position so that only a little influence can
be observed due to the relative displacement
caused by the fault activities.

4500
3500

Shear force (kN)

Case3

Location of
Crashed
Planes

2500
1500
500
-500

200

400

600

-1500

CONCLUSION

-2500

This paper studied about the influence to


shield tunnel structure due to ground slip, crack,
and deformation propagation in sedimentary
layer caused by fault activities. In order to study
the ground displacement effect, linear and nonlinear FEM analyses are performed using joint
elements for crashed planes observed by the
reflection test.
In the case where linear solid elements are
used
for
soil
modeling,
the
relative
displacement is concentrated to the crashed
plane which is modeled by the joint model. On
the contrary, in the case where non-linear solid
elements are used, the displacement due to the
base offset is widely absorbed in the soil.
In this study, the tunnel is existed relatively
very shallow in the ground and the deep deposit
layer is existed below the tunnel. Therefore, the
offset displacement is absorbed in the relatively
bottom part of the soil so that only very little
influence is observed to the tunnel.
Based on the result of this study, it can be
concluded that the influence to the tunnel
structure is very small even though the tunnel
crosses the fault line where the deposit layer is
thick, the tunnel is existed in the shallow depth.

Case2

Shear capacity

-3500
-4500

Distance(m)

Fig. 9: Shear force acted to the tunnel and the


strength
20000

Location of
Crashed
Planes

15000

Moment(kNm)

10000
5000
0
-5000

-10000
-15000

200

400

600

Case3

Moment capacity

Case2

-20000
Distance(m)

Fig. 10: Bending moment acted to the tunnel


and the strength

activities is studied by FEM analysis using


beam element with equivalent tunnel stiffness.
Fig. 9 and 10 show the results of Case 2 and 3,
respectively. The origin of the lateral axis is the
left edge of the modeled area.
Fig. 9 shows that the huge shear force is
observed around the crashed plane where the
base offset is applied. Not only Case2 but also
Case3 shows the same result. However the
shear capacity of the joint used for this segment
is big enough compared to the shear force.
Fig.10 show that the bending moment is
also big compared around the crashed plane
where the base offset is applied. As for Case2,
the maximum moment is observed at the right
side of the offset plane. On the contrary, the

REFERENCES
Adachi Y., Yoshimura S., and Nakata T. (2003)
Influence to underground structure by ground
displacement, Proc. of 12th Japan earthquake
engineering symposium, JEES (In Japanese)
Azuma H. and Kiyomiya O. (2001) Reliability
evaluation of sea bed pipeline suffered from
base offset caused by fault, Proc. of 2nd
improvement of seismic disaster prevention
symposium by failure process analysis, JSCE (In
Japanese)
Bray J.D., Seed R.B. and Seed H.B. (1994):
Analysis of earthquake fault rupture propagation

129

through cohesive soil, Journal of geotechnical


engineering, ASCE, 120,3, pp562-580

Sato Y., Abe K., Okada Y., Simazaki K., and Suzuki
Y. (1989) Relation between earthquake
magnitude and fault parameters, Fault
parameters handbook of Japan, Vol.2 2nd
chapter pp82-92, Kajima books

Cole, D. A.., Jr., and Lade, P. V. (1984) Influence


zones in alluvium over dip-slip faults, Journal of
geotechnical engineering, ASCE, Proc. Paper
18788, Vol. 110, No. GT5, pp599-615

Takada S., Nemat H., and Fukuda K. (2001)


Propose of simplified design method for buried
pipes crossing earthquake fault, Journal of
structural engineering, JSCE, No.668, -54,
pp187-194 (In Japanese)

Donald L., et al.(1994) New empirical relationships


among magnitude, rupture length, rupture width,
rupture area, and surface displacement, Bulletin
of the seismological society of America, Vol.84,
No.4, pp974-1002

Takemura M. (1998) Scaling rule of crust type


earthquake in Japan, Earthquake, Vol.2 No.51,
pp221-228, SSJ (In Japanese)

Freeman T., GILLON M., BERRYMAN K.,


MORIWAKI Y., SOMERVILLE P. and MEJIA L.
(2000) Matahina dam-fault surface displacement
design criteria, CD-ROM of the twelfth world
conference on earthquake engineering, Auckland
NZ

Tani K. and Ueda K. (1991) Location and shape of


discontinuous plane in sandy layer by base fault
displacement, Proc. of 26th geotechnical
engineers meeting,
pp1185-1188 JSCE (In
Japanese)

Irikura K. and Miyake H. (2001) Prediction of strong


motion based on scenario earthquake,
Earthquake, Vol. 110No. 6849-875, SSJ (In
Japanese)

Taniyama T. and Watanabe H. (1998) Deformation


propagation in sandy surface layer caused by
reverse fault activities, Journal of Geotechnical
Engineering, JSCE, 591/ -43, pp313-325 (In
Japanese)

Kennedy, R.P., Chow, A.W. and William, R.A. (1977)


Fault movement effects on buried oil pipeline,
Transportation engineering J., ASCE, Vol. 103,
No.TES, pp617-633
Konagai K. and Johanson J. (2001) Lagrangian
particles for modeling large soil deformations,
Proc. of a workshop on seismic fault-induced
failures, pp99-106
Matsuda T. (1975) Magnitude and return period of
the earthquakes caused by active faults,
Earthquake, Vol.2 No.28, pp269-283, SSJ (In
Japansese)
Newmark N. W. and Hall W. J. (1975) Pipeline
design to resist large fault displacement,
Proceedings of the U.S. National Conference on
Earthquake Engineering-1975, EERI, pp416-425
Onizuka N., Hakuno N., Iwashita K., and Suzuki T.
(1999) Stress and strain propagation due to
base displacement caused by reverse fault
activities, Journal of Applied Science, JSCE, 2,
pp533-542 (In Japanese)
Ramancharla P. K. and Megro K. (2001) Non-linear
numerical modeling of dip-slip faults for studying
ground surface deformation, Proc. of 2nd
improvement of seismic disaster prevention
symposium by failure process analysis, JSCE
Ramancharla P. K. and Megro K. (2001) Applied
simulation of non-linear behavior of dip-slip faults
for studying ground surface deformation, Proc.
of a workshop on seismic fault-induced failures,
JSPS, pp.109-114

130

Shallow Foundations and SoilStructure


Interaction

Study on Applicability of Seismic-isolation Foundation


to Railway Structures
X. Luo1, T. Miyamoto1, T. Imamura1
1

Railway Technical Research Institute, Japan

Abstract
Since seismic-isolation foundations generate relatively large displacements adversely
affecting the running safety of train, it is an important task to assess this influence while
design railway structures with seismic-isolation foundations. In this study, bridges and
viaducts are taken as the objects for assessment of influence due to different types of pile
head connections upon running safety of train during earthquakes. As a result, the
influence of seismic-isolation foundation upon running safety of train is well grasped through
structural dynamic analysis, vehicle running simulation and Running-Safety Assessment
(RSA) based on Spectral Intensity (SI).

INTRODUCTION

After the Hyogoken-Nunbu Earthquake of


Jan. 17, 1995, a new code, Seismic Design
Code for Railway Structures (the Railway Code,
in Japanese) (Railway Technical Research
Institute, 1999) was published. Some new
thought for seismic design have been adopted
by drawing the lesson of the earthquake.
According to the code, the ultimate bending
capacity of pile should be assured during the
intensive Level 2 (L2) earthquakes, which
makes the increase of cross-section and rebar
of the piles. To solve this problem and make an
economical
design,
seismic-isolation
foundations adopted in design is an effective
method.
Though the seismic-isolation
foundations can reduce the action by
earthquake vibration to a structure, on the other
hand, the energy absorption causes big
response displacement of the structure
adversely effecting the running safety of train.
Therefore, how to assess the degree of this
influence becomes a critical problem for design
of railway structures with seismic-isolation
foundations.
In this study, bridges and viaducts are taken
as the objects for assessment of the influence
due to different types of pile head connections
(rigid connection, semi-rigid connection and

131

pinned connection) upon running safety of train


during earthquakes. As a result, the influence
has been well grasped through structural
dynamic analysis, vehicle running simulation
and running-safety evaluation based on
Spectral Intensity (SI).
PUSHOVER ANALYSIS

To evaluate the seismic performance of the


structures with different pile head connections,
a static, non-linear procedure called pushover
analysis was applied to Shinkansen structures.
The structures with rigid pile head connection
are shown in Fig.1 (bridge, Case No.B1), and
Fig.2 (viaduct, Case No.V1). With regard to the
cross sections, the pile length for the bridge is
19.5m (Fig.1) and the pile length for the viaduct
is 25.5m (Fig.2). The surface ground is mainly
constituted by sandy soil and clay. Based on
these cross sections, the pile head connections
were modified with semi-rigid connection and
pinned connection through an optimization
design. As a result, the diameters of the piles
are decreased 10% to 15% compared with the
cases of rigid connection, which proves the
effects of seismic-isolation. The combination of
the structures and the pile head connections,
the decreased diameters are shown in Table 1.
According to the Railway Code (Railway

11700

2454

Seismic coefficient (kh)

Surface
layer

Sandy
soil

Sandy
soil

Clay

-20

3250
3250 1500
9500

-10

Clay

-20

Sandy
soil
Gravel

Cast-in-place RC pile (5 staggered


piles), =13001500

Depth (m)

Depth (m)

-10

21500

19500
1500

Surface
layer

7054

4600

2000

2500

Pulling-out side
pile head yielding

500

4500

2500

50

N Value

0.4

Response
displ. 360mm
Response displ. 435mm
Pulling-out side pile middle yielding

0.2

No.B2(Semi-rigid),Teq=0.765s

No.B3(Pinned),Teq=0.998s

500

Surface
layer

Clay

Clay

Depth (m)
Sandy
soil

-20

300

400

Response displ. 470.4mm


Pullingout side
pile head
yielding

0.6

-10

-10

200

500

Fig.3: Load-displacement relationship of bridge

Seismic coefficient (kh)

Surface
layer

Depth (m)

800

8000

900

100

Displacement at crest () (mm)

2000
3800

0.6

0.7

25500 1400 500


26900

800

0.8

0.0

11300

4500

Pier base yielding

Vs (m/s)

Reaching limit
of seismic
performance II

No.B1(Rigid),Teq=0.695s

Sandy
soil
Gravel

Fig.1: Schematic illustration of bridge with ground


condition

900

Response displ. 297.5mm

1.0

Sandy
soil

-20

0.5
0.4

Response displ. 430mm


Upper part
of column
yielding

Upper
part of
column
yielding

Reaching limit
of seismic
performance II

Response displ. 639.4mm


Pulling-out side pile
middle yielding
Upper part of column & Pulling-out
side pile head yielding

0.3
0.2

No.V3(Pinned),Teq=1.494s
No.V2(Semi-rigid),Teq=1.143s

0.1

No.V1(Rigid),Teq=1.081s

0.0

5400
Cast-in-place RC pile (one column
for one pile), =10001100

-30

Gravel

50

N Value

-30

Gravel

500

Vs (m/s)

Pile head
connection

No.B1
Rigid
No.B2 Bridge
Semi-rigid
No.B3
Pinned
No.V1
Rigid
Viaduct
No.V2
Semi-rigid
(Rigid frame)
No.V3
Pinned

200

300

400

500 600
Displacement at crest () (mm)

700

deformation characteristics of semi-rigid pile


head connection used in the analysis were
based on the reference (Kouda et al, 2005), in
which the concrete strength for the pile head is
equivalently increased due to raised restrain
effects of tense hoops distributed within the pile
head. For Case No.B3 and Case No.V3, the
deformation characteristics of pinned pile head
connection were based on a proposed model
(Tazoh et al, 2001), in which the loaddisplacement relationship depends on the
friction factor of metal.
In Fig.3, it is found that the yield seismic
coefficients of the cases with semi-rigid and
pinned pile head connections are about 10% to
20% lower than the case with rigid connection.
Also,
the
yield
seismic
coefficients
corresponding to the semi-rigid and pinned pile
head connections are larger than the seismic
coefficient 0.35 (G3 site classification) due to
response of Level 1 (L1) design earthquake
motion, which means the structures have

Table 1: Combination of structures and pile head


connections

Structure
type

100

Fig.4: Load-displacement relationship of viaduct

Fig.2 : Schematic illustration of viaduct with ground


condition

Case

Diameter
of pile (mm)
1500
1300
1300
1100
1000
1000

Technical Research Institute), the equivalent


natural periods of the structures should be
calculated by using the pushover analysis
method.
Given in Fig.3 and Fig.4 are the calculated
relationships between the seismic load and the
displacement at crest of the structures shown in
Table 1. For Case No.B2 and Case No.V2, the

132

Rigid pile head connection


Semi-rigid pile head
connection
Pined pile head connection

Response Velocity
Sv (mm/s)

Teq : Equivalent
natural period
of structure

SI =

Car body

Lateral stopper

Calculation
of limit SI

Wheel
Truck

Damper

Rail

2.5

S (h,T )dT
0.1 v
Period (sec)

0.1

70mm
Position for
judgment of
derailment

Spring

2.5

Calculation of
response SI

Spectral Intensity (SI) (mm)

Train running simulation

Response acceleration at
crest of bridge

10000
8000

Rigid

Semi-rigid

Pinned

Semi-rigid

Pinned

6000
4000

Rigid

2000
0
0.5

0.6

0.7

0.8

0.9

1.0

1.1

1.2

Equivalent natural period (Teq) (sec)

Bedrock motion

Fig.5: Procedure for influence assessment of seismic-isolation foundation upon running safety of train

heavy casualty. Therefore, RSA of vehicle


must be conducted in design of railway
structures, which is a unique and complicated
problem in the field of railway dynamics.
With regard to the complicated response of
the vehicles, the index SI is proved suitable for
the RSA based on the concept of energy
balance between the input motions and the
response of vehicles (Luo, 2002, 2005). Given
in Fig.5 is a procedure for RSA with the index SI
against transverse vibration caused by
earthquakes.
Firstly,
the
response
accelerations at the crest of the structures were
calculated by inputting the bedrock motions as
shown in Table 2. In the calculation the soil-pile
interaction was taken into account. Secondly,
the response spectra of velocity were
calculated and the response SI values were
obtained by integrating the velocity from 0.1sec
to 2.5sec. Moreover, the running safety limits
of SI were calculated by vehicle running
simulations. Finally, the RSA of the vehicles
was carried out by comparing the response SI

enough capacity satisfying the seismic


performance I against L1 earthquake.
Compared with the equivalent natural period of
structure of Case No.B1 (rigid, Teq=0.695s), the
period Teq for Case No.B2 (semi-rigid)
increases about 10%, and that for Case No.B3
(pinned) increases about 40%.
The characteristics of the load-displacement
relationship of viaducts shown in Fig.4 are
similar to those of bridges shown in Fig.3.
However, the yield seismic coefficients of the
viaducts are about 20% to 30% smaller than the
corresponding cases of bridges because of
lower stiffness for the viaducts. Also, the yield
seismic coefficients corresponding to the semirigid and pinned pile head connections are
larger than the seismic coefficient 0.4 (G5 site
classification) due to response of L1 design
earthquake motion. Therefore, the structures
designed satisfy the seismic performance
criteria according to the Railway Code. Since
the soft surface ground and the rather low
structural stiffness, the equivalent natural
periods of the viaducts have increased about
50% compared to the bridges with same type of
pile head connection.

Table 2: Equivalent natural period and input


bedrock motions
Equivalent natural Design earthquake
Case
period (Teq )Sec motion on bedrock
No.B1
0.695
L1
No.B2
0.765
1.5 times of L1
No.B3
0.998
0.5 times of Spc.II
No.V1
1.081
Spc.I (L2)
No.V2
1.143
Spc.II (L2)
No.V3
1.494

ASSESSMENT OF RUNNING SAFETY OF TRAIN

Under intense ground shaking, running


railway vehicles might be rolled to induce a
large displacement resulting in derailment
and/or overturning which occasionally causes

133

Spectral Intensity (SI ) (mm)

10000

No.B2
8000

No.V1
No.B3

No.B1

in Fig.7 and Fig.8. The Figures show that the


running safety of the seismic-isolation
foundation can be assured against L1
earthquake, but there may be problems for
earthquakes over LI.

L1

No.V2

No.V3

Spc.II
(L2)
Spc.I
(L2)

6000

1.5 times
of L1
0.5 times
of Spc.II

4000

Limit SI stipulated in the Code

3000

CONCLUSIONS

0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6

The good cost performance of the non-rigid


pile
head connections
(seismic-isolation
foundations) has been proven and their RSA
can be conducted as the same as the common
rigid ones.

Equivalent natural period (Teq ) (sec)

Fig.6: Comparison of limit SI between those


obtained in this study and those stipulated in
the Displacement Limited Code

Spectral Intensity (SI) (mm)

with the limit SI.


Given in Fig.6 are the calculated results of
limit SI corresponding to the cases in Table 2.
Moreover, for comparison a line of limit SI for
common structures stipulated in a code called
Displacement
Limited
Code
(Railway
Technical Research Institute, 2006) was plotted
together.
This limit line was defined by
enveloping the limits (gray lines) for common
structures based on 11 representative
earthquake motions.
From Fig.6 it is
understood that even though the seismicisolation foundations generate quite large
displacement, the values of limit SI against
seismic vibration are higher than that for
common structures stipulated in the code.
Therefore, when designing foundations with
non-rigid pile head connections, the RSA can
be conducted in the same way as that for
common rigid ones. The results of RSA against
L1 and over L1 earthquake motions are shown

REFERENCES
Kouda, M., Hamada, Y., Sando, T., and Aoki, H.
(2005) Modeling Deformation Characteristics of
RC Pile Head with Tense Hoops, Proceedings
of the 60th Annual Meeting of Japan Society of
Civil Engineers, pp.961-962 (in Japanese)
Luo, X. (2002) A Code-Type Provision for Running
Safety Assessment of Train Undergone
Earthquake Motions, Proc., 12th European
Conference on Earthquake Engineering, CD
Version, Paper Reference 462, London, U.K.
Luo, X. (2005) Study on Methodology for Running
Safety Assessment of Trains in Seismic Design
of Railway Structures, Journal Soil Dynamics
and Earthquake Engineering; Vol. 25, No.2,
Elsevier Science Ltd., pp.79-91
Railway Technical Research Institute (1999)
Seismic Design Code for Railway Structures,
published by MARUZEN (in Japanese)
Railway Technical Research Institute (2006)
Displacement Limited Code for Railway
Structures, published by MARUZEN (in
Japanese)
Tazoh, T., Ohtsuki, A., Aoki, T., Mano, H., Isoda, K.,
Iwamoto, N., Ishihara, T., Ohkawa, M. (2001)
Developing a New Method of Pile Head
Connection for Decreasing Construction Cost
and
Increasing
Seismic
Performance,
Proceedings of the 26th JSCE Earthquake
Engineering
Symposium,
pp.881-884
(in
Japanese)

10000
8000

Limit SI stipulated in the Code

6000
4000
2000

No.B1 No.B2

No.V2
No.V3
No.V1
Response SI
No.B3

0
0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6

Equivalent natural period (Teq) (sec)


Spectral Intensity (SI) (mm)

Fig.7: RSA against L1 design earthquake motion


10000
8000
6000

Limit SI
stipulated in
the Code

Limit SI
stipulated in
the Code

4000
2000
0
0.5

1.0

Teq (sec)

1.5 0.5

1.0

1.5 0.5

Teq (sec)

Limit SI
stipulated in
the Code
1.0
1.5 0.5

Teq (sec)

Limit SI
stipulated in
the Code
1.0
1.5

Teq (sec)

(a) 1.5 times of L1, (b) 0.5 times of Spc.II, (c) Spc.I (L2), (d) Spc.II (L2)
Fig.8: RSA against earthquake motions over L1

134

A Study on Dynamic Soil-structure Interaction Effect Based on


Microtremor Measurement of Building and Surrounding Ground Surface
M. IIBA 1, M. Watakabe2, A. Fujii3, S. Koyama1, S. Sakai4 and M. Yasui2
1

Building Research Institute, Japan


2
Toda Corporation, Japan
3
Konoike Construction, Japan
4
Hazama Corporation, Japan

Abstract
Soil structure interaction (SSI) effects are investigated based on a microtremor measurement. The
instruments were set in a 7-storied residential building and on the ground to evaluate the sway and rocking
vibrations. The building constructed by HPC prefabricated method has a flamed structure in longitudinal
direction and a walled structure in transverse direction. Through transfer functions of buildings, predominant
frequencies under SSI model and those under based-fixed condition are calculated. The SSI effect is
remarkable in transverse direction due to predominant rocking effect. Through the random decrement
technique, the damping factor of buildings is obtained. It is founded that the damping factors are around 5 to
7% under the microtremor level.

INTRODUCTION

negligible (Iiba et al., 2002).


For
the
performance-based
design,
behaviors of buildings are clarified more
precisely based on data of earthquake motion
observation. But in general, the earthquake
motion observation is not popular, especially,
for purpose of the SSI phenomena.
A simplified method for incorporating SSI
effects and for calculating spring constants is
presented, to evaluate predominant frequency
of buildings with SSI. The SSI effects of
building are investigated based on a
microtremor measurement on a residential
building (Iiba et al., 2004). Through transfer
functions of the building, predominant
frequencies are calculated.
To investigate the applicability of the
simplified method, results by the simplified
method are compared with those by
microtremor measurement.

Phenomena that foundation displacements


occur due to base shear and overturning
moment is called soil structure interaction
(SSI). The SSI phenomena are the inertial soil
structure interaction. On the other hand, the
effect of SSI related to seismic input motions to
the buildings is called the kinematic soil
structure interaction. As a result of the
displacement of foundations due to the inertial
SSI, characteristics of superstructure are
changed as follows;
a) Elongation of natural period (compared
with base fixed condition)
b) Change of damping factor (compared with
base fixed condition)
The Building Standard Law in Japan and its
related enforcement and notices were revised
for the direction to the performance-based
design from 1998 (Midorikawa et al., 2000).
The calculation method of response and limit
strength was provided for checking structural
serviceability and safety of buildings. In the
calculation, SSI effects should be considered
when the interaction effect would be not

EVALUATION OF PREDOMINANT PERIOD OF


SSI SYSTEM

The
SSI
model
consisting
of
a
superstructure with a mass, a base and sway

135

and rocking springs is illustrated in Fig. 1.


Through an assumption that the influence of
foundation mass and moment of inertia at each
floor on response of SSI system is negligible
(Bielak, 1976), an external force acting to
springs of superstructure, sway and rocking is
only the inertial force of superstructure, as
shown in Fig. 2. In the case that the springs are
arranged in a series, an equivalent spring
constant (Ke) for the SSI system can be
obtained as follows (AIJ, 1996);
H2
1
1
1
(1)
=
+
+
Ke Kb K h K r
Where Kb, kh and Kr are spring constants for a
superstructure, sway motion and rocking
motion, respectively. Circular frequencies are
obtained by following equations;
K
K
Kr
2
(2)
b = b , h 2 = h , r 2 =
m
m
mH 2
Where m is the mass of superstructure, and b,
h
and
r are
circular
frequencies
corresponding to spring constants of each
motion. Based on the relationship of Te=2/e,
Tb=2 /b, Th=2 /h and Tr=2 /r, the
predominant period of the SSI system is
obtained as follows;
Te = Tb + Th + Tr
2

In the same way, an equivalent damping


factor of the SSI system (he) is estimated by
following equation.
3

hb =

(5b)

K r = K r + iK r ' = K r (1 + i 2hr ' ) = K r + icr (6b)


The
impedance
has
a
frequency
dependency. Considering the convenience to
incorporating into design, the value at rest
(frequency is zero) will be used to obtain spring
constant.
(7)
Spring constant: K = K ( = 0 )
To calculate static spring constants for
multiple layers of ground, a cone model is
applied, as drawn in Fig. 3(Wolf, 1994 and Iiba
et al., 2002). The spring constants (Khb, Krb) for

Kb
Cb
Ch

Kr

1 cb
2b m

For the sway and rocking motions, dynamic


impedances with a complex are expressed as
follows;
K h = K h + iK h ' = K h (1 + i 2hh ' ) = K h + ich (6a)

(3)

Kh

Cr

Imaginary Summit

Sway
and Rocking Model
Fig. 1: Sway-rocking model

Radius of Foundation :ro

Ground Surface
Shear Modulus

Building
Rocking Displ. Displ. ub

G1

uh

Z0
Z1

G2

Z2

ur
Sway Displ.

EVALUATION OF SPRING CONSTANTS FOR


SPREAD FOUNDATION

T
T
T
(4)
he = hb b + hh h + hr r
Te
Te
Te
Where hb, hh and hr are equivalent damping
factors. The hh and hr are calculated by
equation (5a). The hb is expressed by cb which
is the viscous damping coefficient of the
superstructure.

K '
(5a)
h = sin 0.5 tan 1
K

Inertial Force

Gi

Zi

Sum. of Displacement

Zn-2

= uh + ur + ub
=F/Kh + FH2/Kr + F/Kb
=F (1/Kh + H2/Kr + 1/Kb)

Gn-1
Gn

Zn-1
Engineering Bedrock

Fig. 3: Cone model and ground condition

Fig. 2: Force vs. displacement in simple SSI system

136

stiffness and Poissons ratio, and distance from


the cone summit for rocking motion of i-th layer
of ground. And rr0 is the equivalent radius for
rocking spring constant.
The soil property is expressed as follows;
(20)
Gi ' = Gi (1 + i 2hi )
Where hi is the damping factor of soil. The
Khb and Krb in Equations (8) and (9) is complex,
when using the complex shear modulus of soils
described
in
Equation
(20).
Dynamic
impedances are dependent on a type of
foundations, their dimension and properties of
soil. As to nonlinearity of the soil, there are two
kinds of nonlinearity. One is based on
nonlinearity of soil when seismic wave comes
up in the soil deposits. The shear stiffness and
damping factor are dependent on the shear
strain of the soil. The other is the nonlinearity
related to contact face between foundation or
pile and soil, so called, local nonlinearity. In the
calculation, only the former, that is, nonlinearity
of the soil property is considered.

sway and rocking motions at bottom for spread


foundation are expressed as follows;
K hb = h K1hb

(8)

K rb = r K1rb
1
h = n
1

i=1
hi
8G1rh 0
K1hb =
2 1

(9)
(10)

(11)

Z Z

hi hi 1
hi = i
(
G
Z
Z
1 h 0 hi Z hi 1 )

i = 1,2, , n 1

G Z

hn = n hn 1
G1 Z h 0
1
r = n
1

i =1
ri

(12a)
(12b)
(13)

K 1rb =

8 G1 rr 0
3 (1 1 )

E Z

(14)
4

Z Z

EVALUATION OF SPRING CONSTANTS FOR


PILE FOUNDATION

r 0 ri
ri = i ri 1
3
3
E
Z
Z
Z
1 r 0 ri 1 ri Z ri 1

i = 1,2, , n 1

As it is conformed that the sway spring


constant of a pile foundation is almost the same
as that of the spread foundation which has the
same dimension of plan configuration (Iiba et
al., 2002), the horizontal spring constant at pile
head is calculated by equation (8). On the other
hand, the rocking spring constant of the pile
foundation is remarkably larger than that of the
spread foundation.
The vertical spring of the pile consists of
springs of friction on pile surface and end
bearing at the pile tip. The spring constant of
friction per unit length (S) is estimated by
following equation (Randolf and Wroth, 1978).

(15a)

E Z
rn = n rn 1
E1 Z r 0
2 1 ,
9
2
Z h 0 = rh 0
Z r 0 = 1 1 rr 0
8
16
Ei = 2(1 + i )Gi

rh 0 = B D
rr 0 = 4

B D
3

(15b)
(16)
(17)
(18)

(19)

Where h is a modification factor for sway


and k1hb is the spring constant for rigid
foundation on semi-infinite uniform layer with
soil property of 1st layer. The 1 is a Poissons
ratio of ground under the foundation. Gi and Zhi
are the shear modulus of i-th layer of ground
and the distance of lower boundary from the
cone summit of i-th layer. And rh0, B and D are
an equivalent radius for the sway spring
constant, length and width of foundation. And r
is the modification factor for rocking and k1rb are
the rocking spring constant for rigid foundation
with semi-infinite uniform layer with soil
property of 1st layer. Ei, i and Zri are the elastic

Sv =

2Ge
log e (2rm / B)

(21)

rm = 2.5 L(1 e )
Ge =

(22)
n

1
1
Gi di e = i di

L i=1
L i=1

(23)

Where L and B= (2Ro) are pile length and


diameter, and Ge and e are an averaged shear
stiffness and Poissons ratio along the pile
length. The di is a height of i-th layer.
The spring constant at pile tip (kb) is

137

The building, which is constructed by HPC


prefabricated method, is composed of girders of
reinforced concrete (RC) structure, columns of
steel RC structure and floors and walls of precast RC boards. The building is supported by
the individual foundation with pre-stressed
concrete piles of 0.35m in diameter. Another
same building is constructed with distance of
about 70 m from the building.
The soil condition at the site is drawn in Fig.
5. The sandy and clayey layers are laminated.
The microtremors in the building and on the
ground surface are measured. Periods of
measurement are 600 or 500 s and 7 or 8 sets
of these periods are recorded. The interval of
records is 0.005 s.
Measuring points are shown in Fig. 4. Three
horizontal and two vertical sensors on the roof
and 1st floor, one horizontal one on the 4th
floor are set in the building. Three horizontal
sensors are installed on the ground surface
with 13m at both sides and 26m at one side far
from the building. The sensor arrangements are
the same in both buildings.
Types of sensors are velocity transducers
with servo type and those with moving-coil type
are used. The unit of measured data is velocity.
The vibration system with surrounding
grounds is assumed to be the sway and rocking
model shown in Fig. 6. Following systems are
considered, based on Fig.6 (Stewart and
Fenves, 1998).

estimated by following;
3 Gb Ro
(24)
kb =
8 1 b
Where Gb and b are the shear stiffness and
Poissons ratio of the engineering bedrock,
respectively. To combine two spring constants,
the vertical stiffness of pile is obtained (Masuda
et al., 1993).
E p Ap (1 e 2 L ) + k b (1 + e 2 L )
(25)
K v = E p Ap
E p Ap (1 + e 2 L ) + k b (1 e 2 L )
Where there is 2 (=Sv/EpAp), and Ep and Ap are
an elastic modulus and a cross-sectional area
of pile.
The rocking spring constant of a pile group
is expressed through the summation of all pile
(m). The rocking spring constants for x and y
axes are as follows.
m

i =1

i =1

K R x = K v y i 2 K R y = K v xi 2

(26)

MICROTREMOR MEASUREMENT

A plan at ordinary floor and a section in the


transverse direction of a building are drawn in
Fig. 4. The building, which has 7-storied
residential one with central corridor at
longitudinal direction, is constructed in a central
area of Tsukuba, Japan. There are three and
nine spans in transverse and longitudinal
directions, respectively. A ratio of transverse to
longitudinal width is 1:3.2 and ratios of height to
width in both directions are 1.47 and 0.64.

V alue

D epth
(m )
4,600

4,600

4,600

4,600

4,750

4,600

4,600

4,600

S oil
type

M ark

10

20

30

40

50

4,600

5,400

L o am
C lay

2,100

6.6

S and

5,400

11.9
13.0

10

C lay

15

20

RF

S and
25

7F

2,70

2,775

Plan

27.6
29.1

5F

34.0
35.4
36.1

2,70

Sensors

2,700

6F

Section

Vertical

30

C lay

35

S an d
C lay
w ith
S an d

3F

2,700 2,700

Horizontal

2,70

4F

C lay

S an d
w ith
C lay

40

41.5

2F

45

G ravel

1F

50.4

50

Fig. 5: Ground condition with soil types and


standard penetration values (N-values)

Fig. 4: Plan and section of building and


arrangement of sensors

138

System

Input

Output
(response)

SRB
ug
ug+uf +H+uB
RB
ug+uf
B
ug+uf +H
ug : Free Surface Ground Motion
uf: Input Loss due to Foundation
uf : Sway
uf: 1st Floor Motion Relative to Free Ground
(herein assumption to uf = uf + uf )
uB: Building Response Relative to 1st Floor
H: Equivalent Height
: Rocking angle at 1st Floor
1
2
3

u
u

H
uBB
H u

HH

uuu
f'
g +u

Surface Ground

Rigid Foundation
without Mass

uu
g

uuf' uuf''

Fig. 6: Displacement of ground surface, sway, rocking and building, and definition of transfer functions

Amplitude

13
12
11
10
9
8
7
6
5
4
3
2
1
0

-building
Transverse

2.87H z

3.25H z
4.14H z

A-building
Longitudinal

Amplitude

13
12
11
10
9
8
7
6
5
4
3
2
1
0

3.35H z
2.63H z

3.33H z

Frequency(Hz)
180

180
A-building
Transverse

135
90
45
0
-45

-90

-135

A-building
Longitudinal

135

Phase Delay (deg.)

Phase Delay (deg.)

Frequency(Hz)

90
45
0
-45

-90

-135

-180

-180
1

Frequency(Hz)
A-building
Transverse

1.0

Frequency(Hz)
A-building
Longitudinal

1.0

Coherence

0.8

Coherence

0.8
0.6

0.6

0.4

0.4

0.2

0.2

0.0

0.0
1

Frequency(Hz)

Frequency(Hz)

Fig. 7: Transfer functions and coherence of SRB-, RB- and B-systems

139

a) Sway and rocking model (SBR-system;


uT/ug)
b) Rocking model (RB-system; uT/(ug + uf) )
c) Building model with base fixed (B-system;
uT/(ug + uf + ur))
Where uT is a total response at top of
building including the ground (ug), sway (uf)
rocking (ur) and building responses (uB). ur is
product of the rocking angle at 1st floor and
height of building.
In case of focusing on the 1st mode, the
equivalent height of building should be used in
calculation of the transfer functions (Naito, et al.,
2003). In this analysis, the total height of
building is used.
Transfer functions and coherence of SRB-,
RB- and B-systems of the A-building are shown
in Fig. 7. The transfer functions are obtained as
follows;
Fourier spectral ratios with 80s in time
length and 40s in running time are calculated.
The number of running Fourier spectral ratios
are 84 and 99 in transverse and longitudinal
directions, respectively. After the average of the
running Fourier spectral ratios, the transfer
functions are obtained with the hanning
windows of 8 times.
The fundamental predominant frequencies
of SRB-, RB- and B-systems in transverse
directions are 2.87, 3.25 and 4.14Hz,
respectively. The reduction of predominant
frequencies occurs when including the sway
and
rocking
motions.
The
amplitude
(amplification factor) at predominant frequency
of SRB-system is the largest, while the
amplitude of B-system is lowest.
By the half power technique, the damping
factors of SRB-, RB- and B-systems are 5.37,
9.24 and 7.63%, respectively.
As to phase characteristics, the phase of
building top in SRB-, and B-systems at the
predominant frequency are faster (the same
result is shown in B-system of longitudinal
direction). Through the detail check for
phase characteristics in SRB-system, about
half of them are faster and the others are
delay. The phase characteristics are not
constant. The average phases are fast or
delay occasionally. In case of RB-system,
the delay cases are more than fast ones.
In longitudinal direction, the fundamental
predominant frequencies of SRB-, RB- and
B-systems are 2.63, 3.33 and 3.35Hz,

respectively. The amplitudes at predominant


frequencies of each system are almost similar.
The damping factors of SRB-, RB- and Bsystems are 8.69, 11.0 and 5.61%, respectively.
As to the coherences, RB- and B-systems
shows relatively large values. In the SRBsystem, since the coherence is less in
frequency range higher than the fundamental
predominant frequency, the input and output
are less corresponding to each other.
The predominant frequencies in various
systems in both directions are summarized in
Table 2. The sway and rocking frequencies are
calculated by using following equation (AlJ,
1996);
2
2
2
1 / f SRB = (1 / f B ) + (1 / f Sway ) + (1 / f Rocking ) (27)
The ratios of predominant frequencies in
SSI system (fSRB) to those at base-fixed
condition (fB) are 0.68-0.69 and 0.75-0.79 in
transverse
and
longitudinal
directions,
respectively. In the transverse direction, ratios
of sway and rocking are 0.22-0.23 and 0.300.31, respectively. In the longitudinal direction,
ratios of sway and rocking are 0.26-0.38 and
0.08-0.18, respectively.
VERIFICATION OF SIMPLIFIED METHOD

Table 3 presents properties of the building.


And equivalent values which contribute to 1st
mode vibration are shown. The spring
constants calculated by Equations 8, 9 and 26
are shown in Table 4. The results of
predominant frequency of the building with SSI
in the transverse direction are also presented in
the Table. After the group of 4 pre-stressed
piles is replaced to a cast-in place RC pile of
Table 2: Predominant frequencies and
ratios to frequency at B-system
Build
ing

A
Transverse
B
A
Longitudinal
B

140

Predominant frequency
of each system (Hz)
Sway

Rocking

4.14

6.12

5.24

(1.00)

(1.48)

(1.27)

3.19

4.11

5.83

5.04

(0.68)

(0.78)

(1.00)

(1.42)

(1.23)

2.63

3.33

3.35

4.29

30.5

(0.79)

(0.99)

(1.00)

(1.25)

(9.10)

2.76

3.21

3.70

5.39

6.47

(0.75)

(0.87)

(1.00)

(1.46)

(1.75)

SRB

RB

2.87

3.25

(0.69)

(0.79)

2.80

0.73m in diameter, the rocking spring is


calculated. The predominant frequency of
building (2.93Hz) is obtained by Equation 3, is
very similar to that in the microtremor
measurement (2.80-2.87Hz) shown in Table 2.
The displacement distribution of the building in
the analysis has a good agreement to that in
the measurement. It is proved that the
simplified method gives appropriate results.

the fundamental predominant frequencies are


calculated by RD (Random Decrement)
technique (Tamura, et al., 1993).
The waveforms of the top of building are
conducted with band-pass filter of 2-4Hz
remained. The waveforms of 5s, when the time
at peak values is adjusted to t=0, are
superposed along 600s (A-building) or 500s (BTable 4: Properties of soil and foundation, and
predominant frequency by simple analysis

DAMPING FACTOR IN SSI SYSTEM

To investigate the damping factor of building,


Table 3: Properties of building and its simple model
Items
Number of story
Dimensions of plan
Total height
Mass of unit area
Mass of each floor
Stiffness of each story
Natural frequency with base
fixed
Equivalent mass at first mode
Equivalent Stiffness at one
degree of freedom model
Equivalent height at first mode

Unit

Values

m
m
t/m2
t
MN/m
Hz

7
41.6 x 12.9
19.0
1.2
6.30x103
205
4.1

t
MN/m

3.60 x103
2.39 x103

15.2

Amplitude(cm/sec)

B-building Transverse
0.000

-0.002
0.5

1.0

1.5

2.0

Time(s)

Values
13.1m for sway spring
9.86m for rocking spring
2.50m
Unit mass
(t/ m3)
1.6

Shear wave
Velocity(m/s)
151

13.0

1.8

260

27.6

1.8

332

41.5

1.7

309

1.9

428

0.002

0.0

Items
Equivalent radius of
foundation
Embedment of
Foundation
Soil
Layer
Depth
Prop(m)
erty
1
6.6

Sway spring constant


Rocking spring constant
Predominant frequency
under SSI without piles
Predominant frequency
under SSI with rocking
fixed
Rocking spring constant
with RC piles (0.73m in
diameter)
Predominant frequency
under SSI with RC piles
Mode
Distribution
ratio
Analysis
Experiment

Fig. 8: Free vibration waveform by RD technique

5.39x103 MN/m
2.59x105 MNm/rad
2.17Hz (0.529: ratio to
base fixed condition)
3.41Hz (0.832: ratio to
base fixed condition)
1.07x106 MNm/rad

2.93Hz (0.714: ratio to


base fixed condition)
Sway Rocking Building
0.226
0.263
0.510
0.210.300.460.23
0.31
0.48

Table 5: Damping factor by RD technique


Direction

Transverse

Longitudinal

Set No.
ATM01
ATM02
ATM03
ATM04
ATM05
ATM06
Ave.
ATL01
ATL02
ATL03
ATL04
ATL05
ATL06
Ave.

A-Building
B-Building
Predominant Damping
Predominant Damping
Number of
Number of
Factor Set No.
Factor
Frequency
Frequency
Superposition
Superposition
(Hz)
(%)
(Hz)
(%)
1801
2.75
6.51
BTM01
1472
3.00
5.81
1796
2.95
5.97
BTM02
1478
2.75
5.04
1779
2.85
5.82
BTM03
1485
2.90
6.17
1771
2.85
5.73
BTM04
1460
2.75
5.04
1777
2.85
5.10
BTM05
1469
2.80
5.96
1796
2.90
6.67
BTM06
1468
2.80
5.32
1787
2.86
5.97
Ave.
1472
2.83
5.56
1785
2.80
5.57
BTL01
1472
3.00
7.48
1772
2.90
5.07
BTL02
1524
2.80
7.75
1792
2.80
7.40
BTL03
1519
2.80
7.45
1833
2.95
7.27
BTL04
1520
2.80
5.99
1833
2.85
7.51
BTL05
1505
2.90
6.16
1823
3.10
6.46
BTL06
1495
2.90
6.31
1806
2.90
6.55
Ave.
1515
2.87
6.86

141

Bielak, J. (1976) Modal analysis for buildingsoil interaction, Journals of Engineering


Mechanical Division, ASCE, Vol. 102, EM5,
pp.771-786
Iiba, M. et al. (2002) Introduction of Soil
Structure Interaction to Performance-based
Type of Provisions in Building Standard Law
of Japan, Proceedings of International
Conference on Advances and New
Challenges in Earthquake Engineering
Research, Harbin, part 2, pp.275-282
Iiba, M. et al. (2004) A Study on Dynamic SoilStructure Interaction Effect Based on
Microtremor Measurement of Building and
Surrounding Ground Surface, Proceeding
of Third US-Japan workshop on SSI, CDROM
Masuda A. et al. (1993) "Simplified Method for
Dynamic
Impedance
of
Pile-group,
Summaries of Technical Papers of Annual
Meeting, AIJ, 339-344 (in Japanese)
Midorikawa, M. et al. (2000) Development of
Seismic
Performance
Evaluation
Procedures in Building Code of Japan,
Proceedings of 12th World Conference of
Earthquake Engineering, Auckland, Paper
No. 2215
Naito, Y., Tanino, T., et al. (2003) System
Identification of Building Alone with the
Effect of Dynamic Soil Structure Interaction
Eliminated Journal of Structural and
Construction EngineeringNo.564pp.3946 (in Japanese)
Randolf M. and Wroth C.P. (1978) "Analysis of
Deformation of Vertically Loaded Piles,
Journals of Geotechnical Engineering
Division, ASCE, Vol.104, No.GT12, 14651488
Stewart, J. and Fenves, G. (1998) System
Identification for Evaluating Soil-structure
Interaction Effects in Buildings from Strong
Motion Recordings, Journals of Earthquake
Engineering and Structural Dynamics, 27,
pp.869-885
Tamura, Y., Sasaki, A. and Tsukagoshi, H.
(1993) Evaluation of Damping Ratios of
Randomly Excited Buildings Using the
Random Decrement Technique Journal
of Structural and Construction Engineering
No.454pp.29-38 (in Japanese)
Wolf J. (1994) Foundation Vibration Analysis
Using Simple Physical Models, PTR
Prentice Hall

building). Obtained waveforms are normalized


to the maximum amplitude at t=0. The
logarithmic damping factors are calculated by
applying the least square technique. An
example of free vibration waveform and
identified curve are drawn in Fig. 8. The
damping factors are summarized in Table 5.
The numbers of superposition are about 1500.
The damping factors in SSI in transverse and
longitudinal directions are 5.0-6.7 and 5.1-7.8%,
respectively. The average damping factors in
each direction are 5.77 and 6.71%. The
damping factor in longitudinal direction is larger
than that in transverse direction.
CONCLUSIONS

The SSI characteristics of building by


microtremor and the verification of simplified
method are summarized as follows;
1) For the performance-based design, a
simplified method incorporating SSI effects is
necessary to obtain precise response of
buildings
2) From microtremor measurement, the ratios
of predominant frequency of SSI systems to
that of building with base-fixed condition are
0.68 - 0.70, in the transverse direction due to
sway and rocking effects.
3) The predominant frequency of building
based on the simplified procedure is very
similar to that in the measurement. It is proved
that the simplified method is effective to
evaluate predominant frequency of SSI system.
4) From microtremor measurement, the
damping factors in SSI systems are evaluated
to be 5 to 7%, using the random decrement
technique.
ACKNOWLEDGEMENTS

The microtremor measurement is conducted


under the research theme entitled Seismic
observation of building and surrounding
ground in the Tsukuba Research Institute
Council. The authors express their sincere
thanks the members of the research group for
their help of microtremor measurements.
REFERENCES

Architectural Institute of Japan (1996) An


Introduction to Dynamic Soil-Structure
Interactionpp.17-19 (in Japanese)

142

Damage Index Evaluations of 3DOF System with Dynamic


Soil-Structure Interaction
K. Kawano1 , Y. Kimura2, Y.Nakamura3
1

Kagohisma University, Japan


Kagoshima University, Japan
3
Kagoshima University, Japan
2

Abstract
For the reliable performance-based design, it is necessary to evaluate the damage of
the structure. The seismic performance evaluations of dynamic soil-structure
interaction system by means of the damage index are examined. The soil-structure
interaction is represented with the sway-rocking model and the nonlinear
characteristics on the structure expressed with the tri-linear model. It is suggested
that it is very important to evaluate the relations of the dominated frequency between
the soil-structure interaction systems and the seismic input motions for the damage
evaluations of soil-structure interaction system due to severe seismic motions.

INTRODUCTION

significant effects on the seismic response


evaluation of the structures.
In the present study, the damage
evaluations by means of combining the ductility
ratio with the seismic response energy to the
SSI system are examined. The soil structure
interaction is represented with the sway-rocking
model (SR Model) and the nonlinear
characteristics on the structure expressed with
the tri-linear model. It is understood that the
ductility ratio and the seismic response energy
are closely related with the damage evaluation
of the structure. If strength characteristics of the
structure have uncertainty, it is expected to
cause the important effect on the damage
evaluations. In order to enhance the damage
evaluation, it is important to verify the
uncertainty effects on the nonlinear seismic
responses.

If some damages can be allowed in the


structural members, it provides very available
methods for the earthquake-proof designs
because of reducing the intensity of the input
forces. From such as background, a lot of
researchers have been carried out for the
performance-based designs. The strength
demand spectrum based on the ductility ratio is
one of the most useful methods which can be
treated with nonlinear effects on the structure
subjected to seismic forces (Iemura et al. 1998).
Recently, the strength demand spectrum based
on the damage evaluation of the structure has
been examined (Mikami et al. 1998). It has
been indicated that the total energy of the
seismic motion on a structure may be given by
one of the important factors on the damage
evaluations (Park et al. 1995).
Moreover, in order to perform the reliable
performance based design, it is important to
clarify for the damage evaluations not only the
dynamic characteristics of structure but also the
dynamic soil-structure interaction effects on the
nonlinear response situation. For the structure
with soil foundation system, it is supposed that
the dynamic soil-structure interaction (SSI) has

FORMULATION

Many researchers have examined for the


dynamic SSI and supposed that the SSI plays
important roles on the seismic response
evaluations. While the finite element method is
one of the most comprehensive methods, it
requires a lot of computing time and does not

143

seem to provide the relevant evaluation for a


preliminary design of the structure. Since the
dynamic responses of the superstructure
subjected to seismic forces mainly depend on
the
dominant
natural
frequency,
the
superstructure is represented with a singledegree-of-freedom (SDOF) system. The
nonlinear characteristics on the structure can
be expressed with the tri-linear model. The
response evaluation of the total system may be
represented with a simplified model as shown in
Figure 1. The governing equation of motion for
the nonlinear SSI system is expressed with

[M ]{&x&}+ [C ]{x&}+ [K (t )]{x} = {F }

k
H

kh

ch

c
VS1
J 0 VS2

m0

kr

cr

Fig.1: An idealized structure-soil-foundation model

(1)

The iterative procedure can be carried out using


the Newton Raphson method.
For damage evaluations of structure
subjected to seismic forces, it is necessary to
evaluate the estimation of energy response due
to seismic motions. Taking into accounts for the
energy on the SSI system, it can be expressed
with integration as follows:

{x&} [M ]{&x&}dt + {x&} [C ]{x&}dt +


t

{x&} [K (t )]{x}dt = {x&} {F }dt


t

(6)

Namely, this equation of motion energy can be


represented as follows:

EK + ED + EH = E

(7)

in which EK , E D , E H and E denote the


kinematic energy, the viscous damping energy,
the hysteretic energy and the total energy of
input seismic motion, respectively. The
hysteretic energy is caused by the restoring
force, and plays significant roles on the energy
evaluation with respect to the nonlinear
response due to seismic motions.
The damage evaluation is conveniently
conducted using the drift displacement and the
energy ratio of the hysteretic energy to the total
input seismic energy. The drift displacement is
depended upon the seismic motion properties
as well as the hysteretic characteristics and
structural properties. For the damage
evaluations by means of combining the ductility
ratio with the seismic response energy, the
damage index to the RC structure by Park et al

(2)

in which

~
~
[ K (t )]{x} = {F }

x1

m1

in which [M] and [C] denote the mass and


damping matrix, and [K (t )] denotes the
stiffness matrix on each time step including the
soil foundation system with linear properties. {F}
denotes vector for the input seismic motion.
The seismic motions corresponding to the
design spectra as shown in Figure 2 are near
field seismic motions recorded in Japan. { &x& },
{ x& } and { x } denote the acceleration responses,
the velocity responses and the displacement
responses, respectively. If the dynamic
response is carried out within linear region, the
governing equation can be solved with modal
analysis and spectral analysis. On the other
hand, if the response of the structure causes to
be nonlinear, it is hard to solve the equation in
the frequency domain. Thus, applying the
incremental method for the equation (1), this
equation can be solved with the increment
method because of the nonlinearity due to the
structure. The equation (1) can be expressed
with the incremental method as follows:

[ M ]{&x&} + [C ]{x&} + [ K (t )]{x} = {F }

x0 H

Zg

(3)

4
2
~
(4)
[ K (t )] = [ K (t )] + ( 2 )[ M ] + ( )[C ]
t
t
4
~
~
~
(5)
{F} = {F} + [M ]{( 2 ){x&}+ 2{&x&}} + [C ](2{x&})
t
Therefore, the increment of the responses can
be determined by solving the equation (3).

144

is applicable to assess the damage situation by


means of the maximum displacement and the
hysteretic energy as follows:

dE

Maximum Acceleration (m/sec**2)

D = 1M +
x1 u
Q y x1u

Kobens
Takans
Portns

30

(8)

in which x1M and x1u stand for the maximum


displacement and ultimate displacement,
respectively. The ultimate displacement can be
determined by the corresponding allowable
ductility ratio. The allowable ductility ratio is 5.0
in this present study. Qy denotes the yield force

25

20

15

10

0
0.1

Natural Period (sec)

of the structure and denotes the coefficient


depending on the characteristics of the
structural member (Fajfar 1992). The damage
index is practically related to the damage
situations by means of the damage assessment
of real structures experienced to seismic forces
as shown in Table 1. In this present study, the
strength demand spectra based on the damage
index from 0.1 to 1.0 are examined.

Fig.2: Acceleration response spectra

Table 1 :The relationship between Damage


Index and Damage level
Damage Index

Damage level

00.1

Slightly

0.10.2

Light

0.20.4

Moderate

0.41.0

Severe

1.0

Failure

RESULTS AND DISCUSSIONS


In order to clarify the damage evaluation of
the SSI system as shown in Figure 1, the
seismic response analysis with the SSI system
is carried out using typical seismic motions
corresponding to the design spectra as shown
in Figure 2. In this present study, the soilfoundation is supported by ground condition
which has two layers and the sway and rocking
spring
constant of soil-foundation can be
determined to the each ground condition
(Yamada et al. 1979). Table 2 shows the shear
wave velocity of each ground condition due to
the seismic motion as shown in Figure 2. The
structure has the viscous damping ratio 0.05,
and the nonlinearity of the structure expressed
with the tri-linear model. The seismic response
analysis is implemented with the Newmarks
method and the time increment is used to be
0.001 sec.
Figure 3 shows the comparison of natural
period of the superstructure with SSI system. It
is understood that the natural period of the SSI
system is longer than the superstructure in the
relatively high frequency range due to the
dynamic soil-structure effects.

Table 2: The shear wave velocity of each ground


condition for seismic motion
K obens

T ak a n s

P o rtn s

300

150

100

350

300

300

(Unit:m/sec)

On the other hand, there are few differences in


the low frequency range. The natural period of
SSI system is supposed to be mainly depended
on the superstructure. From this result, it is very
important to clarify the effects of dynamic soilstructure interaction for the damage evaluations
of the structure.

145

Demand Acceleration (m/sec**2)

Narural Period of the SSI system (sec)

35

Kobens
VS1=300m/sec,VS2=350m/sec
Takans
VS1=150m/sec,VS2=300m/sec
1

Portns
VS1=100m/sec,VS2=300m/sec

0.1
0.1

30
25
20
15
10
5
0

Natural period of the superstructure (sec)

0.1

Fig.3: Comparison the natural period of


superstructure with SSI system

Natural Period of the superstructure (sec)

Fig.4 : A strength demand spectra (Kobens)

demand spectra on damage


35

Demand Acceleration (m/sec**2)

Strength
index

SDOF
D=0.1
SSI system
D=0.1
D=0.2
D=0.4
D=0.6
D=1.0

Input seismic motion:Kobens


Damping ratio:0.05

The strength demand spectra are used to be


examined with the ductility ratio. However, the
damage evaluations can not be always carried
out using only the strength demand spectra
based on the ductility ratio. Therefore, it is
necessary to examine with other method of
assessment for the damage evaluations. The
strength demand spectra based on the damage
index from 0.1 to 1.0 are examined in this
present study.
Figure 4 and Figure 5 show the strength
demand spectra of the SSI system subjected to
the Kobens and Takans as shown in Figure 2,
respectively. The abscissa denotes the natural
period of the superstructure, and the ordinate
denotes
the
demand
maximum
input
acceleration. Each line corresponds to the
damage index from 0.1 to 1.0. The strength
demand spectrum of SDOF is shown by
comparison with the SSI system in each figure.
Comparing the demand acceleration of an input
seismic motion for the damage index, very
effective reduction of demand acceleration can
be caused by increase of the damage index.
For the input seismic motion, Kobens, the
strength demand spectra has a significant
difference between the SDOF and the SSI
system because of the dynamic SSI effects in
the relatively high frequency range. On the
other hand, since the strength demand spectra
due to the Takans yield relatively smaller
difference than the Kobens, it is understood that
the effects of dynamic soil-structure interaction

30

SDOF
D=0.1
SSI system
D=0.1
D=0.2
D=0.4
D=0.6
D=1.0

Input seismic motion:Takans


Damping ratio:0.05

25
20
15
10
5
0
0.1

Natural Period of the superstructure (sec)

Fig.5 : A strength demand spectra (Takans)

become smaller than the Kobens.


As previously mentioned, the strength demand
spectra are evaluated to the allowable ductility
ratio, 5.0. It is necessary to examine the
ductility ratio corresponding to the assigned
damage ratio for the SSI system. Figure 6
shows the maximum ductility ratio of SSI
system subjected to the Kobens. The abscissa
denotes the natural period of the superstructure,
and the ordinate denotes the maximum ductility
ratio of the SSI system. Each line corresponds
to the damage index from 0.2 to 1.0. For the
maximum ductility ratio due to the damage
index,1.0, it is understood that the maximum
ductility ratio has become under an allowable
ductility ratio for the natural period of the
superstructure from 0.2 sec to 2.0 sec. However,
for the maximum ductility ratio to the Kobens, it
is understood that the maximum ductility ratio
exceeds an allowable ductility ratio in the high

146

frequency range. For the maximum ductility


ratio to the Takans, the maximum ductility ratio
exceeds an allowable ductility ratio in the low
frequency range.
For the frequency region such as exceeding
the allowable ductility ratio, it is supposed to be
reexamined of the target damage index. A
constant damage index is obtained in case the
damage level of the structure is slight. Namely,
the damage index depends on the natural
period of the structure as well as the dynamic
characteristics of the input seismic motion.
However, if the severe damage level of
structure can be allowed, it is understood that
the damage index causes to some variations for
the natural period of the superstructure. The
damage index, which can be expressed in
Equation (8), is expressed with combining the
ductility ratio and the hysteretic energy of the
structure. When the damage level of the
structure is severe, the hysteretic energy also
increases, and it is indicated to be considerable
effects on the damage index. If the
performance- based design is applied to the
design of low damage level situation, the
maximum ductility ratio of the structure could be
evaluated with the damage spectra. However,
for the severe damage level of the structure, it
is necessary to carry out the damage
evaluations by means of combining the ductility
ratio with the hysteretic energy

7
6

Maximum ductility ratio

SSI system
D=0.2
D=0.4
D=0.6
D=1.0

Input seismic motion:Kobens


Damping ratio:0.05

5
4
3
2
1
0
0.1

Natural Period of the superstructure (sec)

Fig.6 : Maximum ductility ratio (Kobens)

7
6

Maximum ductility ratio

SSI system
D=0.2
D=0.4
D=0.6
D=1.0

Input seismic motion:Takans


Damping ratio:0.05

5
4
3
2
1
0
0.1

Natural Period of the superstructure (sec)

Fig.7 : Maximum ductility ratio (Takans)

Figure 8 shows the ratio of hysteretic energy


to total energy of SSI system subjected to
Kobens. The assigned damage index is
assumed to be 0.4 and 1.0. The abscissa
denotes the natural period of the superstructure,
and the ordinate denotes the ratio of hysteretic
energy to total energy. It is understood that the
ratio of energy gradually increased in the high
frequency range of superstructure by the
influence of the dynamic SSI. Comparing the
ratio of energy of SDOF with the SSI system, it
is understood that the ratio of hysteretic energy
lead to some differences under the natural
period of the superstructure, 0.5 sec. For the
SDOF system, the ratio of energy on the
damage index D=0.4 is about 60% and D=1.0 is
about 70%. For the SSI system, the ratio of
energy on the damage

147

100

Iuput seismic motion:Kobens


Damping ratio:0.05

90

SDOF
D=0.4
D=1.0
SSI system
D=0.4
D=1.0

80
70

EH/E (%)

Evaluations due to energy response

60
50
40
30
20
10
0
0.1

Natural Period of the superstructure(sec)

Fig.8 : Energy ratio due to the damageindex


(Kobens)

index D=0.4 is about 40% and is about 60% for


the damage index D=1.0. From these results, it
is understood that the SSI plays important roles
on the energy evaluations. In order to enhance
the estimation of the damage of the structure, it
is necessary to clarify the influence of the
dynamic SSI which is mainly depended on the
dynamic characteristics of the input seismic
motion and the natural period of the structure.

Drift displ. / Yield displ.

1.5

Input seismic motion:Portns


Damping ratio:0.05
D=0.2
D=0.4
D=0.6
D=1.0

1.0

0.5

Evaluations due to drift displacement


0.0
0.1

It is indicated that the drift displacement could


be a very important value from the viewpoint of
restoration at the early time of the structure
(Kawashima et al. 1994). The ratios of the drift
displacement to the yield displacement can be
represented with the drift displacement spectra
because it is evaluated with the demand
strength spectra by means of changing the yield
displacement in this present study. Figure 9
shows the drift displacement spectra to the
Portns. The abscissa denotes the natural
period of the superstructure, and the ordinate
denotes the ratio of drift displacement to yield
displacement. For damage index D=0.2, it is
noted that the ratio of the drift displacement to
the yield displacement indicates relatively small
value. It is noted that the ratio of the drift
displacement to the yield displacement
becomes approximately constant values in spite
of the natural period of the superstructure.
However, as the damage index increases, the
ratio leads to be very large fluctuation and it is
primarily depended on the natural period of the
superstructure. Moreover, it is understood that
the drift displacement yields large fluctuation for
not only the natural period of the superstructure
but also the damage index. Especially, the drift
displacement spectra corresponding to damage
index D=1.0 have very remarkable fluctuation.

Natural Period of the superstructure (sec)

Fig.9: Drift displacement spectra (Portns)

Demand Acceleration (m/sec**2)

15
Input seismic motion:Kobens
Damping ratio:0.05

D=0.4
D=0.6

12

0
0.1

Natural Period of the superstructure (sec)


Fig.10: A strength demand spectra with uncertainty (Kobens)

Demand Acceleration (m/sec**2)

15

Influence of the structural uncertainty


In general, the strength properties of the
structure are expec ted to have various
uncertain factors. Since the uncertainty could
yield the important contributions on the damage
evaluations, it is essential for the performancebased design of the structure to clarify the
uncertainty effects with respect to the nonlinear
response. The damage index shown in

Input seismic motion:Portns


Damping ratio:0.05

D=0.4
D=0.6

12

0
0.1

Natural Period of the superstructure (sec)


Fig.11 A strength demand spectra with uncertainty (Portns)

148

equation (8) is related to the yield displacement,


the maximum displacement, the expected
ultimate displacement and hysteretic energy.
Taking into account for the random process of
the input seismic motion and the structure
properties, it is important for the damage
evaluation of the structure to clarify the effects
of the uncertainties. In the present study, the
strength demand spectra with respect to the
uncertainty of the yield displacement are
examined. The coefficient of variation is
assumed to be 20%. Figure 10 shows the
strength demand spectra of SSI system with
uncertainty of yield displacement to Kobens.
Figure 11 similarly shows the strength demand
spectra of SSI system with uncertainty of yield
displacement to Portns. The abscissa denotes
the natural period of the superstructure, and the
ordinate denotes the demand acceleration. A
solid line and dotted line denote the strength
demand spectra due to the damage index
D=0.4 and D=0.6 in case of without uncertainty,
respectively. A dashed line and dashed-dotted
line denote the strength demand spectra to
uncertainty, respectively. Each response is
presented together with the one corresponding
to the mean value plus (or minus) the standard
deviation. Comparing the damage index, 0.4,
with the one 0.6, the demand acceleration is
mainly affected by extent of the uncertainty. If
the structure has the uncertainty of yield
displacement, it could be supposed to lead to
the severe damage level instead of the
moderate damage level as described by Table
1.From the results, it is important for the
damage evaluation to verify the influence of the
uncertainty.
.

to clarify the dynamic soil-interaction effects.


(2) The damage evaluation shows the different
tendency with respect to a maximum
response ductility ratio according to the
allowable
damage
level.
For
the
performance-based design, the light level of
damage can be evaluated by the maximum
ductility ratio of the structure. However, for
the severe damage level of the structure, it
is necessary to carry out using the damage
evaluations by means of combining the
ductility ratio with the hysteretic energy.
(3) The ratios of the drift displacement to the
yields displacement have a tendency
fluctuated by the natural period of the
superstructure. Increasing the damage
index, it leads to the fluctuation for the
natural period of the superstructure. Since
the drift displacement gradually increases
for the increment damage index, it is
important to carry out the available
evaluation for the drift displacement.
REFERENCES
Iemura,H., Igarashi,A. & Takahashi,Y.( 1998)
Ductility and Strength demand for near field
earthquake ground motion: Comparative study on
the Hyogo-ken Nanbu and the Northridge
earthquakes, Structural safety and Probability
pp.1705-1708
Mikami,T., Hirao,K., Sasada,S., Sawada,T. &
Nariyuki,Y.( 1998), A study on design spectra of
seismic intensity for level 2 earthquake, The 10th
Earthquake Engineering Symposium, pp.30613066 (in Japanese)
Park,Y.-J. & Ang,A.H.-S.(1995), Mechanistic
seismic damage model for reinforced concrete,
Journal of Structural Engineering, Vol.111, No.4,
pp.722-739
Yamada,Y, Takemiya,H.,Kawano,K,( 1979).
Random response analysis of a non-leanear
soil-suspension bridge pier, Earthq. Eng. Struct.
Dyn. Vol.7, pp.31-47
Kawashima,K., Macrae,G.A., Hoshikuma,J. &
Nagaya,K. (1994), Residual displacement
response spectrum and its application, Journal
of structural mechanics and earthquake
engineering, JSCE, No.501/ -29, pp.183-192 (in
Japanese)
Fajfar,P. (1992), Equivalent Ductility Factors,
Taking into Account Low-Cycle Fatigue,
Earthquake Engineering and Structural Dynamics,
Vol.21, No.10,pp.837-848

CONCLUSIONS
The seismic performance evaluations of SSI
system by means of the damage index are
examined. The main results are summarized as
follows:
(1) The damage index combined the maximum
displacement and the hysteretic energy can
be applied to estimate the strength demand
spectra for the SSI system. The damage
evaluation depends on the ground condition
and the dynamic characteristics of input
seismic motions. It is necessary for the
evaluations of the strength demand spectra

149

Seismic response of rigid structures stepping on nonlinear foundation


A. Palmeri1, N. Makris2
1
2

Department of Civil Engineering, University of Messina, Italy


Department of Civil Engineering, University of Patras, Greece

Abstract
The rocking motion of slender/rigid structures stepping on nonlinear yielding foundation is
examined. This work is the continuation of previous investigations on rocking structures
where the foundation behavior was restricted to linear viscoelastic. With yielding supporting
springs, the geometric nonlinearities from the dynamics of a rocking block combine with the
material nonlinearities of the foundation. This paper focuses in assessing the effects of the
geometric and material nonlinearities and identifies various trends of the dynamic response.
Selective results are presented.

a base undergoing horizontal motion were


presented by Housner (1963). Following this
seminal work, a large number of studies have
been presented to address the complex
dynamics of the free-standing block (Zhang &
Makris, 2001, and references provided therein),
while a handful of studies examined the rocking
response of structures stepping on a non-rigid
foundation. Psycharis and Jennings (1983)
presented an equivalent linear analysis of the
rocking response of a rigid block stepping on a
viscoelastic foundation, Yim and Chopra (1984)
presented a linear analysis of a flexible
structure supported on a base allowed to uplift,
while Gazetas (1999) documented the
permanent tilting of buildings due to foundation
failure. More recently, Palmeri and Makris
(2005) investigated in depth the rocking
response of a rigid block which steps on linear

INTRODUCTION

The possibility of tall structures to separate from


their foundation limits the earthquake generated
forces and moments, and in most cases leads
to more economical designs, assuming that
stability prevails. For instance, the piers of the
South Rangitikei Viaduct in New Zealand (Beck
& Skinner, 1974; Fig. 1 left), the Rio Vista
Bridge in Sacramento, California (Yashinsky &
Karshenas, 2003; Fig. 1 centre), and the North
Approach Viaduct of the Lions Gate Bridge in
Vancouver, British Columbia (Dowdell &
Hamersley, 2000; Fig. 1 right) are allowed to
rock on their foundation.
The analysis of the rocking response of a
rigid block on a monolithic foundation is well
known to the literature. Early studies on the
rocking response of a rigid block supported on

Fig. 1: Examples of bridges allowed to rock on their foundation.

150

R0

K,y
A

L
D

Fig. 2: Regimes of motion of a rigid block rocking on yielding foundation.

viscoelastic supports via a fully non-linear


analysis, which complements the linear one
presented by Psycharis and Jennings (1983).
In this paper, we present a study on the
rocking response of a rigid block which steps on
yielding supports. It is assumed that the
foundation system cannot take tension;
therefore, separation occurs when the upward
displacement of the heel of the structure is
greater than the static deflection of the support.
The developed methodology is applied to the
response analysis of a bridge peer that is
allowed to step, with the aim of investigating the
effects that a non-rigid foundations has on the
rocking response of tall/slender structures.

POSITION OF THE PROBLEM

We consider in this study the rigid block of mass


M shown in Fig. 2, in which R0 is the size and
is the slenderness. The nonlinear supports at the
points L and R are yielding springs with vertical
stiffness K = v2 M 2 and yielding deflection y .
The springs do not take tension, and the rigid
block can pivot about points L and R when it is
set to rocking. In Fig. 2, the horizontal dotdashed line is used for the position where the
supporting springs are unstressed.
When the motion is weak, then the body is in
continuous contact with both supports at pivot
points L and R. When the vertical uplift of the
heel of the body exceeds the static deflection of

151

the supporting springs, s , then separation


occurs: e.g., in the stage B of Fig. 2 the
separation is imminent, i.e. the supporting spring
at the heel (point R) is unstressed, while in the
stage C the body is supported only at its toe
(point L); the contact at the heel is successively
re-established in the stage E, where the support
is waiting in the unstressed configuration (see
stages C and D).
According to Fig. 2, the rigid block can
experience three regimes of motion: (i) fullcontact at both points L and R (stages A, B, and
E to G in Fig. 2); (ii) contact only at point L
(stages C and D, in which the block is tilting to
the left); and (iii) contact only at point R (stage H
and I, in which the block is tilting to the right).
There is a fourth regime of motion, when the
rigid body separates from both supports, which is
not the subject of this study: when this happens,
our analysis terminates.
It is worth noting that, given the elastoplastic
behaviour of the supporting springs, Fig. 2
shows the plastic deformations accumulated at
the pivot points (e.g., the transition between
stage C and D), which may determine a
permanent tilting of the block at the end of the
motion. As an example, in the stage E of Fig. 2
the elastic deformations at the pivot points L
and R are the same, as in initial stage A;
however, in the stage E the rigid block shows a
nonzero rotation due to the plastic deformation
cumulated at the pivot point L, while prior to the
stage A the supporting springs have not
experienced plastic deformations.

The loss of energy caused by the impacts is


governed in our analysis by the ratio of the
velocity of the imminent pivot point before the
impact to the velocity of the pivot point after the
impact at the support. This ratio defines the
coefficient of restitution, = v&La v&Lb = v&Ra v&Rb , which
has been introduced by Psycharis and Jennings
(1983). Given the abovementioned parameters,
the rotation of a rigid block rocking on a twosupport yielding foundation can be expressed by:
= f ( , , p, , v ,,excitation )

where = M R02 / I0 is the shape factor of the


rigid block, being I0 the moment of inertia of
the rigid block with respect to the centre of
mass; = y s is the bearing capacity index
of the foundation; and p is a frequency
parameter, which results from its s\ize, R0 , and
the intensity of the surrounding gravitational
field, g :
p=

g
+ 1 R0

(2)

Eq. (1) indicates that there are six


parameters of the rocking structure-foundation
system that influence its response other the
characteristics of the ground excitation.
For a bridge tower, the frequency parameter
can be as low as p = 0.5rad/s or even smaller;
whereas for a household refrigerator p = 2.0
rad/s . The slenderness of a 10 m tall classical
column can be as low as = 10 0.17 rad;

v
O

R0
R
L

v&&g (t )

(1)

u&&g (t )

Fig. 3: Schematic of a rigid block rocking on yielding foundation

152

F
Fy

F,v

n
K

Fig. 4: Rheological model (right) and hysteretic loop (centre) for the ideal elastic-perfectly plastic behaviour.
Hysteretic loop of the Bouc-Wen model (left).

the bearing capacity index, = y s , are in


the range 0 1 and > 1.

whereas, the slenderness of a three-storey


building which is 8 m wide is approximately
= 36 0.62 rad. The vertical frequency, v ,
can originate from the vertical stiffness of an
elastoplastic device or the stiffness of a
conventional foundation, depending on the
configuration of the rocking structure.
Accordingly, practical values of interest for the
ratio = v p are in the range 10 < < 100.
Of course, there may be special cases where
the ratio lies outside this range. Finally, the
values of the coefficient of restitution, , and

EQUATIONS OF MOTION

Since the rigid block is not allowed to slide, the


system has only two degrees of freedom: the
vertical displacement of the centre of mass, v ,
and the rotation of the block, (see Fig. 3).
Using the two degrees of freedom, the equations
of motion in the full-contact regime can be
derived in the form:

&& g
&&
v + 2 ( zL + zR ) = g v g (t )
(3)

&&g (t )
u
1

+
2
2
2
1 + cos ( ) && +
p {[ sin( ) + cos( ) ] zL [ sin( ) cos( ) ] zR } = (1 + m ) p cos( )


2
g

where u&&g (t ) and v&&g (t ) are the time histories of


the horizontal and of the vertical components of
the ground acceleration, respectively; while zL
and zR are the dimensionless hysteretic
variables associated with the yielding springs at
the pivot points L and R, respectively, which are
proportional with the corresponding reaction
forces:

FL = Fy zL ; FR = Fy zR

(4)

in which Fy = K y is the yielding force of the


supporting springs. When the rigid block
separates from the support on the right, the
associated hysteretic variables goes to zero,
zR = 0, and the equations of motion are:

&& g
v + 2 zL = g v&&g (t )

u&& (t )
1 + cos2 ( ) && + 1 + p 2 [ sin( ) + cos( ) ] z = (1 + ) p 2 [cos( ) sin( ) ] g
L

2
g

(5)

In the same way, when the rigid block


separates from the support on the left, zL = 0,
and the equations of motion are:
&& g
v + 2 zR = g v&&g (t )

u&& (t )
1 + cos2 ( ) && + 1 + p 2 [ sin( ) cos( ) ] z = (1 + ) p 2 [cos( ) + sin( ) ] g
R

2
g

153

(6)

Each of Eqs. (3), (5) and (6) rules the


dynamic equilibrium of the rigid block in one of
the three regimes of motion depicted in Fig. 2.
These equations, however, have to be coupled
with those ruling the time variation of the
hysteretic variables zL and zR , whose values
depends in principle on the whole time histories
of the vertical displacements of the pivot points
L and R, respectively:

effectively approximated by the Bouc-Wen


model (Bouc, 1967; Wen, 1976), which is
described by the equations:
F (t ) = K y z(t )
z&(t ) =

where n is a dimensionless parameter that


controls the rate of the transition between the
elastic and the perfect plastic phases (Fig. 4
right). In the following, it is always assumed
n = 5.
Taking into account Eqs. (7), the state
equations ruling the hysteretic variables zL and
zR when the rigid block is in contact with the
supporting springs can be derived in the form:

v L = v + R0 sin( ) ; v R = v R0 sin( ) (7)


Among the rheological models available in
the literature for the hysteresis of yielding
springs, the simplest one is the elastic-perfectly
plastic model (Fig. 4 left and centre), made of
an elastic spring with stiffness K in series with
a Coulomb element with sliding force Fy = K y .
The ideal elastic- perfectly plastic model can be

2 p2
sin( )
z&L =
v& +

g
1+

2 p2
sin( )
&
z
=
v&
R
g
1+

(8)
v&
n 1
n
1 0.5 sign(v& ) z z 0.5 z

g sin( ) &
& 1 0.5 sign v& +
zL
(1 + ) p 2

g sin( ) &
& 1 0.5 sign v&
zR
(1 + ) p 2

n 1

zL 0.5 zL

n 1

zR 0.5 zR

(9)

separates from the left spring zL = 0 and


z&L = 0.

When the block separates from the right


spring, the corresponding hysteretic variables
takes the constant value zR = 0, and then
z&R = 0. In the same way, when the block

Table 1: Information pertinent to the strong motions selected in this study.


Earthquake
(Date)

Station
(Component)

D
km

PGD
m

PGV
m/s

PGA
g

Loma Prieta,
California
(18-Oct-1989)

16 LGCP
(000)

6.9

6.1

0.412

0.948

0.563

Erzikan, Turkey
(13-Mar-1992)

95 Erzikan
(NS)

6.9

2.0

0.273

0.839

0.515

7.1

8.5

0.410

1.274

1.497

6.7

6.4

0.327

0.130

0.843

6.7

7.1

0.288

1.661

0.838

6.9

0.3

0.358

0.127

0.611

Cape Mendocino,
89005 Cape
California
Mendocino
(25-Apr-1992)
(000)
Northridge,
24514 Sylmar Olive
California
View Med FF
(17-Jan-1994)
(360)
Northridge,
77 Rinaldi
California
Receiving Station
(17-Jan-1994)
(228)
Kobe, Japan
(16-Jan-1995)

0 Takatori
(000)

NS: North-South.

154

1992 Erzican
PGA= 0.515 g

0.12

max

1995 Kobe
PGA= 0.611 g

0.12

0.08

0.08

0.04

0.04

0.12

max

1989 Loma Prieta


PGA= 0.563 g

1994 Northridge, Rinaldi


PGA= 0.838 g

1994 Northridge, Sylmar


PGA= 0.843 g

1992 Cape Mendocino


PGA= 1.497 g

0.12

0.08

0.08

0.04

0.04

: Rigid foundation
Case No. 1 (+): =40, =2.5
Case No. 3 (): =80, =2.5

Case No. 2 (,): =40, =3.5


Case No. 4 (*): =80, =3.5

Fig. 6: Maximum rotation values of the bridge tower when subjected to the strong motions listed in Table 1.

can see that for the first three ground motions a


foundation with larger values of stiffness and/or
yielding deflection produces larger responses; in
these cases, moreover, the model with rigid base
(solid lines) gives conservative estimations of the
peak rotation. Remarkably, in two among the last
three ground motions the opposite happens.
Nevertheless, what is important to note is that large
variations in the mechanical properties of the
foundation have just a small effect on the value of
the peak rotation. Furthermore, the tower, although
very slender ( = 0.13 rad), experiences always
small rotations because it has such large size
R0 = 60.1m).
( p = 0.36 rad/s,
and
More
precisely: i) the maximum peak rotation of the
stepping bridge under analysis (max = 0.114 ) is
induced by the 1989 Loma Prieta earthquake
(PGA = 0.563 g ) when the base is assumed rigid;
and ii), contrary to what one would intuitively
expect,
the
minimum
peak
rotation
(max = 0.0191 ) is induced by the 1992 Cape
endocino earthquake, which exhibits the maximum
peak ground acceleration (PGA = 1.497 g ), also
in this case when the base is assumed rigid.

NUMERICAL APPLICATIONS

The proposed nonlinear formulation is used to


estimate the seismic-induced response of a
stepping bridge about 60 m tall to six strong
ground motions chronologically listed in Table 1.
Frequency parameter, angle of slenderness and
shape factor of the tower are assumed to be
p = 0.38 rad/s, = 0.13 rad ( 7.4) and = 3,
respectively. Moreover, the coefficient of restitution
is = 0.05, while four combinations of parameters
and are considered: 1) = 40 and = 2.5;
2) = 40 and = 3.5 (larger yielding deflection);
3) = 80 (stiffer foundation) and = 2.5; and 4)
= 80 and = 3.5. The peak rotations computed
with these combinations of the foundation
parameters are also compared with that one
computed with a monolithic base.
Fig. 6 plots the peak normalized rotations of the
tower of interest when subjected to the six ground
motions listed in Table 1, including also the vertical
component. They are ordered with increasing peak
ground acceleration, from PGA = 0.515 g (1992
Erzikan) to 1.497 g (1992 Cape Mendocino). One

155

= 40, = 2.5
0

&
p

= 40, = 3.5

10 15 20 25 0

10 15 20 25 0

= 80, = 2.5
5

= 80, = 3.5

10 15 20 25 0

10 15 20 25

0.12

0.12

0.06

0.06

-0.06

-0.06

-0.12

-0.12

0.6

0.6

0.3

0.3

-0.3

-0.3

-0.6

E
M

-0.6

-1
-1

u&&g (t )
g

0.6

0.6

0.3

0.3

-0.3

-0.3

-0.6

-0.6
0

10 15 20 25 0

t [s]

10 15 20 25 0

10 15 20 25 0

t [s]

t [s]

10 15 20 25

t [s]

-4 -3 -2 -1 0 -3 -2 -1 0 -4 -3 -2 -1 0 -3 -2 -1 0 -8 -6 -4 -2 0 -6 -4 -2 0 -8 -6 -4 -2 0 -6 -4 -2 0
0

F -0.25
Fy -0.5

-0.5

-0.75
-1

-0.25

-0.75

-1

-4 -3 -2 -1 0 -3 -2 -1 0 -4 -3 -2 -1 0 -3 -2 -1 0 -8 -6 -4 -2 0 -6 -4 -2 0 -8 -6 -4 -2 0 -6 -4 -2 0

v y

v y

v y

v y

Fig. 7: Time histories of rotation , angular velocity & and total energy E (top graphs) of a bridge
pier with parameters = 3, = 0.13 rad and p = 0.38 rad/s rocking on a yielding foundation
with coefficient of restitution = 0.05 and four combinations of parameters and , when
subjected to the 1989 Loma Prieta earthquake. Hysteretic loops v F (bottom graphs) of the
yielding springs..

156

It is worth noting that, although is some cases


the peak rotations computed for different
configurations of the foundation are very close
each other, the time histories of the rocking
response may be quite different. In Fig. 7, for
instance, normalized rotation (first graph from
the top), angular velocity & (p ) (second graph)
and total energy E M (third graph, with values in
J/kg) of the tower of interest when subjected to
the 1989 Loma Prieta earthquake are compared.
When the foundation is rigid (solid lines), the peak
rotation takes the larger value, while the
oscillations looks very regular. In the case of
yielding foundation, on the contrary, the peak
rotations take slightly smaller values, while the
rocking responses prove to be very sensitive to
stiffness and load bearing capacity of the
foundation. For instance, the hysteretic loops of the
nonlinear supporting sprigs (first graphs from the
bottom) show that = 80 and = 2.5 is the case
where the inelastic demand is larger (third column).

ACKNOWLEDGMENTS

We thank the European Social Fund (ESF),


Operational Program for Educational and
Vocational Training II (EPEAEK II), and
particularly the Program PYTHAGORAS II, for
funding the above work.
REFERENCES

Beck, J.L., and Skinner, R.I. (1974). The seismic


response of a reinforced concrete bridge pier
designed to step, Earthquake Engineering and
Structural Dynamics, Vol. 2, pp. 343-58.
Bouc, R. (1967). Forced vibration of mechanical
systems with hysteresis, Proc. 4th Conf. on
Nonlinear Oscillation.
Dowdell, D.J., and Hamersley, B.A. (2000). Lions
Gate Bridge North Approach Seismic retrofit,
Proc. 3rd International Conference STESSA 2000.
Gazetas, G. (1999). Overturning and settlement in
Adapazari during the 1999 Izmit (Turkey)
Earthquake, Proc. 1st International Conf. on the
Kocaeli Earthquake.

CONCLUSIONS

In this paper, the rocking response of


slender/rigid structures stepping on nonlinear
yielding foundation is investigated. The response
analysis of this study complements previous
investigations where the foundation behaviour is
restricted to linear viscoelastic. A novel set of
equations of motion is presented, where two
hysteretic variables are included, aimed to account
for the plastic deformations cumulated in the
supports. In the proposed model the geometric
nonlinearities arising from the dynamics of a
rocking block combine with the material
nonlinearities of the foundation. As a result, the
response of the rocking block to a given
accelerogram depends on its size, shape and
slenderness, together with the stiffness and the
yielding deflection of the foundation, and with the
amount of energy dissipated during the impacts.
The numerical applications show that a simple
model with rigid base may give accurate
estimations of the peak rotation, while more refined
models are required when the inelastic demand in
the foundation is of importance. Some
counterintuitive results are also emphasized: for
instance, an increase in the stiffness and/or in the
strength of the foundation may indifferently
produce larger or smaller rotations.

Housner, G.W. (1963). The behaviour of inverted


pendulum structures during earthquakes, Bulletin
of the Seismological Society of America, Vol. 53, pp.
404-17.
Palmeri, A., and Makris, N. (2005). Response analysis
of rigid structures rocking on viscoelastic
foundation, Report No. EEAM 2005-02,
Department of Civil Engineering, University of
Patras, Greece.
Psycharis, I.N., and Jennings, P.C. (1983) Rocking of
slender rigid bodies allowed to uplift, Earthquake
Engineering and Structural Dynamics, Vol. 11, pp.
57-76.
Wen, Y.K. (1976). Method for random vibration of
hysteretic systems, Journal of Engineering
Mechanics Division ASCE, Vol. 102, pp. 249-63.
Yashinsky, M., and Karshenas, M.J. (2003).
Fundamental of seismic protection for bridges,
EERI Monograph MNO-9, Earthquake Engineering
Research Institute, Oakland, California.
Yim, C.K., Chopra, A.K., and Penzien, J. (1980)
Rocking response of rigid blocks to earthquakes,
Earthquake Engineering and Structural Dynamics,
Vol. 8, pp. 565-87.
Zhang, J., and Makris, N. (2001) Rocking response
and overturning of free-standing blocks under
cycloidal pulses, ASCE Journal of Engineering
Mechanics, Vol. 127, pp. 473-83.

157

Numerical simulations of shaking table experiments on a shallow


foundation test model at PWRI, Japan
R. Paolucci1, M. Shirato2 , M.T. Yilmaz3
1

Politecnico di Milano, ITALY


Public Works Research Institute, Tsukuba, JAPAN
3
Middle East Technical University, Ankara, TURKEY
2

Abstract
The set of experiments carried out at the Public Works Research Institute, Tsukuba, Japan,
is a unique opportunity to analyze the seismic behaviour of shallow foundations under
seismic loading of various levels of intensity up to complete failure. The experimental set up
and the different phases of the experiments are described in more detail in a companion
paper for this Workshop (Shirato et al., 2007). In this contribution the experiments are
simulated using a simplified approach consisting of a single degree-of-freedom structure
founded on a compliant foundation with 3 degrees-of-freedom (vertical, horizontal, rocking).
The nonlinear behaviour of the soil-foundation system is allowed by numerical modelling
using a nonlinear elasto-perfectly plastic macro-element, defined by a suitable yield surface
and plastic flow rule.
The results of the numerical analyses are very satisfactory in terms of the accurate
simulations of the time history of overturning moments, driving the nonlinear behaviour of
the foundation. The simulation of the accumulated settlements and rotations was much
more difficult: satisfactory results were obtained by using a suitable stiffness degradation
rule as a function of the total plastic rotation accumulated during shaking. These results
throw light on the debated issue of the foundation ductile behaviour during strong seismic
shaking and on its possible role in the evolution of performance-based design concepts
including nonlinear soil-structure interaction.

INTRODUCTION

design, there is now an increasing awareness


of the effects of the interaction between the
foundation and the superstructure and its role
on the overall seismic capacity of the system
(see e.g. ATC-40, 1996; Martin and Lam, 2000;
Pecker 2006). This awareness is conflicting with
the lack of reliable methods for predicting
foundation yielding under strong earthquake
shaking and for putting it into proper
consideration in the framework of such
approaches.
For this purpose, nonlinear dynamic finite
element simulations of large numerical models,
including the superstructure, the foundation and
the surrounding soil, are probably not the best
tool, since they require the use of sophisticated
soil constitutive laws and very large
computational
times
to
perform
a
comprehensive set of parametric analyses. To
overcome such limitations, but preserving at the

In the framework of seismic design


according to capacity principles, it is generally
recognized that any damage to the foundation
should be avoided. This means that the
nonlinear capacity of the system is exploited at
the superstructure level alone, typically
permitting the energy dissipation at suitably
selected points through formation of plastic
hinges or insertion of isolation/dissipation
devices. This requirement is partly supported by
budget considerations, due to the time and
economic costs related to the post-earthquake
inspection and verification of the foundation
system, but it is also justified based on the lack
of well established and calibrated methods to
study the post-yielding behavior of soilfoundation systems under strong seismic loads.
With the ever increasing interest towards
performance based approaches for seismic

158

defined according to Nova and Montrasio


(1991) and a plastic flow rule according to
Cremer et al. (2002). Further details on the
numerical procedure are reported in Paolucci et
al. (2007).

same time the essential features of the dynamic


soil-structure interaction problem, the macroelement concept has become more and more
popular (Nova and Montrasio, 1991; Paolucci,
1997; Cremer et al., 2001 and 2002; Le Pape
and Sieffert, 2001). It basically consists of
modelling the soil-foundation system as a single
non-linear macro-element, having 3 degrees of
freedom (dof) for in-plane analyses, defined by
a suitable failure surface and plastic flow rule.
However,
although
the
macro-element
approach is quite promising, it has not been
supported so far by adequate experimental
evidence, at least for seismic applications.
Indeed, few experimental results are
available on the nonlinear soil-shallow
foundation interaction under dynamic seismic
loads, including, among the most relevant ones,
the quasi-static large-scale tests performed at
the Joint Research Center at Ispra, Italy (Negro
et al., 2000; Faccioli et al., 2001) and at the
Public Works Research Institute (PWRI, 2005)
at Tsukuba, Japan, and the fully dynamic tests
in centrifuge (Zeng and Steedman 1998, Gajan
et al., 2005) and in shaking table (Maugeri et al.
2000).
In this work we make reference to the
companion paper of Shirato et al. (2007)
presented at this Workshop, which summarizes
the experimental setup and main results of a
comprehensive set of shaking table tests on the
performance of a model shallow foundation
resting at the surface of a laminar box filled with
dry sand, and excited by real accelerograms of
various levels of amplitude. The purpose of this
contribution is to show that, after a suitable
improvement of the method proposed by
Paolucci (1997), it is possible to capture with a
satisfactory agreement many of the details of
the experimental response, even during the
highly nonlinear phases of foundation behavior.

x0

x1

m1
k1,c1

k0,c0
m0
kr,cr

kv,cv

xg
yg

Fig 1: The 4 degrees-of-freedom model used for


dynamic nonlinear soil-structure interaction
analyses

A further improvement of our results was


obtained by taking into account that during the
strongly nonlinear phase of the excitation the
instantaneous foundation-soil contact area
decreases, owing to successive cycles of
foundation rotations. This was implemented by
a simple degradation rule for the foundation
stiffness parameters (k0, kv and kr, see Fig. 1),
described by the following equation:

B = B (1-D)

(1)

where B is the foundation width and D is a


degradation parameter defined as follows:

( )

D p =

COMPUTATIONAL PROCEDURE

D1
1 + 1 D2 p

(2)

where D1 and D2 are model parameters related


to the ultimate D value and to the degradation
speed, respectively, while p is the cumulated
plastic foundation rotation at a specified instant
of time. The best agreement with observations
has been obtained using D1 = 0.75 and D2 =
5000/rad. Details on this model and on its
implementation in the numerical procedure are
given in Paolucci et al. (2007).

The computational procedure is based on the


simplified approach proposed by Paolucci
(1997), consisting of a two-dimensional soilstructure interaction model having 4 dof (Figure
7), one to account for the horizontal motion of
the superstructure, assumed to behave in the
linear-elastic range, while the soil-foundation
system is represented by a 3 dof (horizontal,
rocking and vertical motion) elasto-perfectly
plastic macro-element, with yielding function

159

NUMERICAL SIMULATIONS RESULTS

without
degradation
is
already
rather
satisfactory, the improvement obtained by the
stiffness degradation rule (1) and (2) leads in
many cases to capture some important details
of the nonlinear response, including the period
elongation following the major yielding phases
of the model response. Note that, while in the
calculations with the non-degrading model the
damping value was kept constant to the 5%
value estimated from the small-amplitude
vibrations, in the degrading model we raised
such value to 10% (15% for Case 1-2), in order
to limit the amplitude of fluctuations of the
computed response, especially in terms of
rotation, and improve the agreement with the
observations. The higher frequency excitation
for Case 1-2, and the corresponding shortduration yielding phases, can explain the higher
damping ratio required to fit the observations.
The parameter set used for the analyses is
reported in Table 1, while the reader is referred
to Paolucci et al. (2007) for further details on
their selection.
When considering foundation rotation (Fig. 4),
the agreement is again quite satisfactory,
especially if we consider the loading magnitude
and the large permanent effects observed on
the test model. Again, the degrading model is
suitable to capture some important details of
the observed time histories, except for Case 1-4
for which the numerical calculation diverges.

We have studied the response of the model to


the four earthquake input motions considered in
the sequence of shaking table experiments
(Fig. 2), that will be denoted in the following as
Case 1-2, 1-4, 1-5 and 2-2, according to the
sequence of tests.

Fig 2: Earthquake records used as base excitations


for shaking table tests: (a) 1993 Hokkaido
Nansei Oki Earthquake Schichihou Bridge
record, (b) and (c) 1995 Kobe Earthquake
JMA record, (d) 80% scaled 1995 Kobe JMA
record

A sample of results is shown in Fig.3 in terms of


the normalized overturning moment M/VB. We
have omitted Case 1-5 for sake of brevity,
mainly because it was run just after Case 1-4,
without any intermediate preparation work of
the foundation soil, so that the dynamic soil
properties are not the same as in the other
experiments. The horizontal displacements will
also be omitted in the discussion, since they
generally result in negligible permanent
deformations. All plots in Fig. 3 show the
normalized overturning moment observed (thick
lines) and simulated (thin lines) within the 10 s
long most severe phase of the excitations.
Within this phase, the foundation behaviour is
highly nonlinear, and frequently reaches a
threshold eccentricity ratio that in our
experiments turns out to be slightly higher than
0.4. For all cases, we have compared both
results with and without stiffness degradation,
to show that, although the simplest approach

Table 1: Elastic parameters used for numerical


simulations
k0
N/m
3.0107
cv
Nms/
rad
1.8104

160

kr
Nm/ra
d
2.0106

kv
N/m

k1
N/m

c0
Ns/m

3.8107

3.0109

1.6104

cr
Nms
/rad
2.0103

c1
Ns/m

m1
kg

m0
kg

h
m

J
kgm2

0.

664

161

0.54

1.5102

Fig 3: Time histories of the normalized overturning moment observed (thick lines) and simulated (thin lines)
either with or without stiffness degradation. The small picture at the bottom shows a zoom of the
observed vs. simulated comparison during the pre-yielding phase of the excitation. From Paolucci et al.
(2007)

161

Fig 4: The same as Fig. 3, in terms of foundation rotation time history. From Paolucci et al. (2007)

Fig 5: The same as Fig. 3, in terms of foundation vertical settlement. From Paolucci et al. (2007)

162

Faccioli E, Paolucci R, Vivero G. (2001)


Investigation of Seismic soil footing interaction
By large scale cyclic tests and analytical
models., 4th Int. Conf. on Recent Advances in
Geotechnical Earthquake Engineering and Soil
Dynamics, Special Presentation Lecture SPL-05,
San Diego.

The main limitation of our numerical


calculations is clearly the vertical settlement
prediction, as shown in Fig. 5: our elastoperfectly plastic macro-element model is not
presently able to simulate the accumulation of
large vertical settlements throughout the
excitation. However, when the magnitude of
settlements and the number of loading cycles is
limited as in Case 2-2, the prediction of the final
vertical displacement is relatively satisfactory,
although details of the time history are missed.
Comparing rotations and vertical displacements
in Figures 4 and 5, it seems that there is a
coupling of these degrees of freedom in the
nonlinear range, that is not suitably accounted
for in our simplified model.

Gajan S, Kutter B, Phalen J, Hutchinson TC, Martin


GR. (2005) Centrifuge modeling of loaddeformation behaviour of rocking shallow
foundations. Soil Dynamics and Earthquake
Engineering, Vol. 25: pp. 773-783.
Le Pape Y, Sieffert JP. (2001). Application of
thermodynamics to the global modelling of
shallow foundations on frictional material.
International Journal for Numerical and Analytical
Methods in Geomechanics, Vol. 25, pp. 13771408.

CONCLUSIONS

Martin GR, Lam IP (2000). Earthquake resistant


design of foundations - Retrofit of existing
foundations; Proc. GeoEng2000, Melbourne,
2000; pp. 1025-1047.

The numerical method used in this study,


based on the macro-element concept to
account for in a simplified way the nonlinear
dynamic
interaction between
soil
and
foundation, has been proven rather successful
to simulate the results of the sequence of
shaking table experiments on a shallow
foundation test model carried out at the Public
Works Research Institute, Tsukuba, Japan. The
agreement
with
observations
is
quite
satisfactory for the rotational degree of
freedom, for which the introduction of a simple
stiffness degradation rule allowed us to capture
also some important details of the foundation
response. Still some progress should be made
as regards the prediction of vertical settlements,
for which a more sophisticated constitutive
model is required.

Maugeri M, Musumeci G, Novit D, Taylor CA.


(2000). Shaking table test of failure of a shallow
foundation subjected to an eccentric load. Soil
Dynamics and Earthquake Engineering, Vol. 20:
pp. 435-444.
Negro P, Paolucci R, Pedretti S, Faccioli E. (2000).
Large-scale
soil-structure
interaction
experiments on sand under cyclic loading, Proc.
12th World Conference on Earthquake
Engineering, Auckland, NewZealand, paper #
1191.
Paolucci R. (1997). Simplified evaluation of
earthquake induced permanent displacements of
shallow foundations Journal of Earthquake
Engineering, Vol. 1, pp. 563-579.
Paolucci R. Shirato M, Yilmaz MT (2007). Seismic
behaviour of shallow foundations: shaking table
experiments vs. numerical modelling. Submitted
for publication to Earthquake Engineering &
Structural Dynamics.

REFERENCES
ATC-40. Seismic evaluation and retrofit of concrete
buildings. Chapter 10: Foundation effects.
Technical Rep. SSC 96-01, Seismic Safety
Commission, State of California, 1996.

Pecker A (2006). Enhanced seismic design of


shallow foundations: example of the Rion
Antirion bridge. 4th Athenian Lecture on
Geotechnical Engineering, Athens, Greece,
January 2006.

Cremer C, Pecker A, Davenne L. (2001) Cyclic


macro-element for soil-structure interaction:
material
and
geometrical
non-linearities.
International Journal for Numerical and Analytical
Methods in Geomechanics,Vol. 25: 1257-1284.

Shirato M, Nakatani S, Fukui J, Paolucci R (2007)


Large-scale model tests on shallow foundations
subjected to earthquake loads. Proc. 2nd
Japan-Greece Workshop on Seismic design,
observation and retrofit of foundations, 3-4 April
2007, Tokyo.

Cremer C, Pecker A, Davenne L. Modelling of


nonlinear dynamic behaviour of a shallow strip
foundation with macro-element. Journal of
Earthquake Engineering, 2002; 6: 175-212.

163

Zeng X, Steedman RS. (1998) Bearing capacity


failure of shallow foundations in earthquakes,
Gotechnique, Vol. 48, pp. 235-256.

164

Analytical Modelling of Footings under Large Overturning Moment


M. Apostolou and G. Gazetas
National Technical University of Athens, Greece

Abstract
Non linear features of the rocking response of tall structures founded on shallow
foundations are investigated. To this extent a macroscopic modelling of the soil
foundation system is developed, capable of representing the large-displacement domain
of the response. Analytical equations for the monotonic load-displacement relationship
are extracted incorporating both geometric and material nonlinearities. Such analytical
backbone curves may be implemented in dynamic SSI analysis through the concept of
nonlinear macro-element to represent the near-field soil-foundation system. The
limitations of conventional Winkler-based modelling are also highlighted under strong
overturning moments.

INTRODUCTION

The use of overstrength factors, is


necessitated by the so-called capacity
design principle, under which plastic hinging
is allowed only in the super-structural
elements not in the below-ground (and thus
un-inspectable)
foundation
and
soil.
Therefore, structural yielding of the footing
and mobilisation of bearing capacity
mechanisms is not allowed. Only a limited
amount of sliding deformation and uplifting at
the foundationsoil interface is allowed.
However, there is a growing awareness in the
profession of the need to consider soil
foundation inelasticity, in analysis and
perhaps even in design [see: Pecker (1998),
Martin & Lam (2000), and Allotey & Naggar
(2003)]. This need has emerged from:

The conventional approach to foundation


design introduces factors of safety against
sliding and exceedance of ultimate capacity,
in a way similar to the traditional static design.
This approach involves two consecutive steps
of structural and foundation analysis:

Dynamic analysis of the structure is


performed in which the soil is modeled as
an elastic medium, represented by
suitable translational and rotational
springs (and, sometimes, with the
associated dashpots).

The dynamic forces and moments


transmitted onto the foundation are
derived from the results of such analyses
along with considerations for inelastic
structural response (e.g. by reducing the
moments in columns through the
behaviour [ductility] factor q). The
foundations are then designed in such a
way that these transmitted horizontal
forces
and
overturning
moments,
increased by overstrength factors, would
not induce sliding or bearing capacity
failure.

165

The large (often huge) acceleration (and


velocity) levels recorded in several
earthquakes which are associated with
even larger elastic spectral accelerations
(of the order of 2g). Enormous ductility
demands would be imposed to structures
by such accelerations if soil and
foundation yielding did not effectively
take place to limit the transmitted
accelerations.

interface and the passivetype resistance


often enjoyed by embedded foundations.

In seismically retrofitting a building or a


bridge, allowing for soil and foundation
yielding is the only rational alternative.
Because increasing the structural capacity
of some elements would imply that the
forces transmitted onto the foundation be
increased, to the point that it would not be
technically or economically feasible to
undertake them elastically. Thus, new
retrofit design guidelines (e.g. FEMA 356)
explicitly permit inelastic deformations in
the foundation.
Even with new structures, it has been
recognised that with improved analysis
methods we need to better evaluate
performance in terms of levels of damage.
For the superstructure, performance
based
design
or
equivalently
displacementbased design have been
used for a number of years, with inelastic
push-over analyses becoming almost
routine in seismic design practice. It is
logical to extend the inelastic analysis to
the supporting foundation and soil.

Sliding at the soil-foundation


interface

Foundation uplifting from the


supporting soil

NEW DESIGN PHILOSOPHY: PLASTIC


HINGING IN SHALLOW FOUNDATIONS

Bearing capacity type of


soil failure

Excluding structural yielding in the isolated


footing or the foundation beam, three types of
nonlinearity (shown schematically in Fig 1)
can take place and modify the overall
structurefoundation response:

Fig 1: Plastic hinge approach at soilfoundation


interface.

Separation and uplifting of the foundation


from the soil. This would happen when the
seismic overturning moment tends to produce
net tensile stresses at the edges of the
foundation (Fig 1b). The ensuing rocking
oscillations in which uplifting takes place
involve primarily geometric nonlinearities, if
the soil is competent enough. There is no
detriment to the vertical load carrying capacity
and the consequences in terms of induced
vertical settlements may be minor. Moreover,
in many cases, footing uplifting is beneficial
for the response of the superstructure, as it
helps reduce the ductility demands on
columns. Housner (1963), Pauley & Priestley
(1992), and many others have reported that
the satisfactory response of some slender
structures in strong shaking can only be
attributed to foundation rocking. Deliberately
designing a bridge foundation to uplift in

Sliding at the soilfoundation interface.


This would happen whenever the transmitted
horizontal force exceeds the frictional
resistance (see Fig 1a). As pointed out by
Newmark (1965), thanks to the oscillatory
nature of earthquake shaking, only short
periods of exceedance usually exist in each
one direction; hence, sliding is not associated
with failure, but with permanent irreversible
deformations. The designer must only ensure
that the magnitude of such deformations
would not be structurally or operationally
detrimental. Although this philosophy has
been applied to the design of earth dams and
gravity
retaining
walls,
its
practical
significance for foundations might be
somewhat limited in view of the large values
of the coefficient of friction at soilfooting

166

analyses in which inelastic action in the soil is


considered or is ignored. With inelastic action
(including uplifting) the shear wall sheds
some of its load onto the columns of the
frame, which must then be properly
reinforced; the opposite is true when linear
soilfoundation behaviour is assumed. Thus,
computing the consequences of plastic
hinging in shallow foundation analysis may
be a necessity. The interplay between uplifting
and mobilization of bearing capacity
mechanisms is governed primarily by the
following factors:

rocking has been proposed as an effective


seismic isolation method by Kawashima &
coworkers (2005). Moreover, even with very
slender and relatively rigid structures, uplifting
would not lead to overturning except in rather
extreme cases of little concern to the engineer
(Makris & Roussos 2000, Gerolymos et al
2005). In soft and moderately-soft soils much
of what was said above is still valid, but
inelastic action in the soil is now unavoidable
under the supporting edge of the uplifting
footing in rocking. At the extreme, inelastic
deformations in the soil take the form of
mobilization of failure mechanisms, as
discussed below.

Mobilization of bearing capacity failure


mechanisms in the supporting soil. Such
inelastic action under seismic loading would
always be accompanied with uplifting of the
foundation (Fig 1c). In static geotechnical
analysis large factors of safety are introduced
to ensure that bearing capacity modes of
failure are not even approached. In
conventional seismic analysis, such as in the
EC8 Part 5 bearing capacity is avoided
thanks to an overstrength factor of about
1.40. The oscillatory nature of seismic
shaking, however, allows the mobilisation (for
a short period of time!) of the maximum soil
resistance along a continuous (failure)
surface. No collapse or overturning failure
occurs, as the applied (causative) moment
quickly reverses, and a similar bearingcapacity failure mechanism may develop
under the other edge of the foundation. The
problem again reduces to computing the
inelastic deformations, which in this case
means permanent rotation. The designer must
ensure that its consequences are not
detrimental.

the vertical foundation load N in


comparison with the ultimate vertical
capacity Nu, expressed through the ratio
= /u
the height, h, of the mass center of gravity
from the base compared with the
foundation dimensions (width B, length L)
the intensity, frequency content and
sequence of pulses of the seismic
excitation.

CONVENTIONAL WINKLER MODELLING

In common SSI analysis procedures,


Winkler modelling of soil has often been found
a simple and efficient tool because of its ability
to incorporate different nonlinear aspects of
the rocking behaviour at relatively low levels
of computational cost. For example, notension springs can capture uplifting effects at
the soil-foundation interface whereas soil
yielding can be represented with elastic
perfectly plastic springs. In the context of a
conventional Winkler model the following
postulations are often encountered:

The concept of allowing mobilization of


bearing capacity mechanisms in foundation
design may represent a major change in
foundation design philosophy (FEMA 1997,
Pecker 1998). However, for analysis of the
ultimate response of a structurefoundation
system to extreme earthquake shaking,
accounting for such a possibility is necessary.
Martin & Lam (2000) illustrate with an
example of a hypothetical structure containing
a shear wall connected with a frame how
dramatically different are the results of

167

A unique constant spring modulus kv is


adopted: for any type of loading
(symmetric or antisymmetric), independent
of the distance from the mid-point x .
Correspondingly, a purely vertical or
moment loading results in a uniform or
triangular distribution of contact pressures
along the foundation.
The rotation pole of the structural system
remains fixed at the foundation mid-point
even after uplift or soil yielding initiates.
During a clockwise (counter-clockwise)
rotation, uplift onsets when the vertical

displacement at the left edge (right edge)


of the foundation becomes zero.
Second order ( P ) effects are ignored,
even in the domain of large lateral
displacements of tall superstructures.

Elastic soil

For purely elastic soil behaviour,


nonlinearity of the momentrotation response
emerges once uplifting occurs and leads to an
ultimate capacity of Mu = Nb . During fullcontact conditions (state 1), the moment with
reference to the centre of the footing can be
easily extracted by integrating distributed
moments over the foundation:

A schematic of a rigid footing resting on


Winkler supports is shown in Fig 2. Bartlet
(1979) used a conventional soil model of
distributed elasticperfectly plastic springs to
study the rocking response of a footing on
clay. His considerations on the different states
of moment-rotation response have been
implemented by the FEMA guidelines
(273/274 documents) in modelling shear walls
as portrayed in the graph of Fig 3. The
following considerations of the Mtheta
relationship are presented in this plot:
(a) the extreme case of elastic soil
conditions represented with the path 1-2-3-5a,
(b) the case of a soil with limited strength
where uplift occurs before yield (path 1-2-3-45b-6), and
(c) the case of a soil with limited strength
where yield precedes the uplifting (path 1-2c3c-4c-5c-6).

M1 =

2
kv x d x =

(=

uplift

b N /(2kv b ) ) together with Eq 1:

uplift =

N
Nb
and Muplift =
= M2
2
2kv b
3

(2)

Upon the onset of uplifting, the moment


rotation curve enters a softening mode (state
3 and 5a). The moment at this level is derived
with integration over the remaining in contact
part of the footing 2 as follows (Siddharthan
et al., 1992):

M3,5 a =

wbo

x2 d x

b 2

kv

pole of
rotation

2b

1 2N
= Nb 1
3 k b 2
v

2N
= Muplift 3

kv b 2

(a)

(1)

The rocking rotation and moment at


incipient uplift (state 2) can be derived by the
uplifting
criterion
w ( b ) = 0

N=mg

p(x)

2kv b 3

po

po (uplift initiation)

(3)

This curve reaches a maximum value of


max M = Nb which represents the ultimate
moment capacity of the foundation for elastic
soil response when P effects are ignored.

po

+
pu po

(b)

(yield initiation)

Elastoplastic soil

Fig 2: Conventional beam-on-winkler-foundation


model: (a) configuration of the model and
(b) superposition of soil contact pressures
at full-contact state.

In the general case of soil with limited


strength, an ultimate value of the spring
reaction pu = Nu / 2b is implemented with Nu

168

The equation of the ultimate moment can


be also obtained from moment equilibrium at
limit state 6 (see Fig 3). After some algebraic
manipulations, Eq 7 yields to the analytical
solution of the failure curve in the N M
plane:

being the foundation bearing capacity under


central vertical load. It is revealed that the
inverse of the safety factor under purely
vertical loading,

= N / Nu = ( FSv )

(4)

N
M = Nb 1

Nu

has a significant effect on the rocking


behaviour (Allotey et al., 2003). In particular,
two separate modes of the moment-rotation
curve can be distinguished dependent upon
the value of : an uplift prevailing state
corresponding to a lightly-loaded foundation
( < 0.5 ) and a soil-yield prevailing state in

(8)

or in nondimensional form
=
m

n
(1 n )
2

= M / Nu B ( B = 2b ) are
n = N / Nu = and m
respectively the nondimensional vertical load
and moment.

case of a heavily-loaded foundation ( > 0.5 ) .


The former follows the path 1-2-3-5b-6 in the
graph of Fig 3 whereas the latter is
represented with the path 1-2c-3c-4c-5c-6. As
shown in Fig 2b uplift initiates before soil yield
if po < pu po or po < 0.5 pu ( < 0.5 ) . The
rocking rotation and moment at uplift onset
are given by Eq 2. Similarly if po > 0.5 pu

(b)
(a)

5a
4

( > 0.5 )

soil yield occurs first. In this case


the yield criterion becomes (Siddharthan et
al., 1992):

5b

5c

(c)
3c

4c

2c

q
N
y = u
kv b 2kv b 2

(5)

and

My =

2qu b 2 Nb

= M 2c
3
3

Allotey et al., (2003) derived the analytical


relationship during concurrent uplift and yield
(state 5b or 5c):

M5 b,c = Nb

pu 3
N2

2 pu 24 ( kv )2

N2
2 pu

3c Yield prior to uplift

2 Elastic at uplift

5c Yield after uplift

3 Elastic with uplift

6 Limit state

(a) Elastic soil behaviour

(6)

(c) Elastoplastic soil behaviour


(yield initiates before uplift)

Fig 3: Rocking of a shear wall on strip footing: the


different states of the M-theta curve under
monotonic loading (Bartlet, 1979, reprinted
in FEMA274 document).

Regardless the value of the ultimate


moment capacity of the foundation derived
from Eq 6 ( for ) is:
M6 = Nb

1 Elastic prior to uplift

The analytical expressions of the

interaction curves in the dimensionless n m


plane for foundation failure (overturn) and

(7)

169

the optimum foundation behaviour in terms of


the moment ultimate capacity is obtained:

incipient uplift and soil yield are summarized


in Table 1 together with their values
at = 0, 0.5, and 1. These curves are also
plotted in Fig 4. The loading plane bounded
by the failure curve is partitioned in regions of
linear and nonlinear response. The nonlinear
area is subdivided in smaller parts depending
on whether uplift or soil yield or both are
encountered. It is worthy of note that a
perfectly symmetric response is achieved
about the vertical axis at = 0.5 . At this point

1
Nu b
4

max Mu = 0.125Nu B =

(9)

In addition, at = 0.5 the maximum range


of linearly elastic response is achieved before
nonlinear conditions due to uplift or soil yield
are engaged, with a threshold moment
of 0.083 Nu B = ( 2 / 3 ) max M .

plane for foundation


Table 1 Analytical expressions of the interaction curves in the dimensionless n m
failure (overturn) and incipient uplift and soil yield.
Interaction curve
Vertical load
factor:
=

(1)
Foundation
uplift:

N
Nu

uplift =
m

Muplift

y =
m

Nu B

0.5

My

Mu
Nu B

=
m

Nu B

rigid =
m

Mu ,rigid
Nu B

(3 4 )
6
1
(1 )
6

6
1
(1 )( 4 1)
6

0.5

(4)
Overturn on
rigid soil:

(3)
Failure
(overturn):

(2)
Soil yield:

(1 )

=0

= 0.5

1
12

1
12

1
8

=1

1
8
1
2

0.25
(1)
(4)

Mu : Nu B

0.2
0.15

(3)
(2)

(d)

0.1

(1)
(3)

(b)

0.05

(c)

((2)
4)

(a)
0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

N : Nu
(a)

Elastic,
full-contact

(b)

Elastic,
uplift

(c)

Yield,
full-contact

(d)

Yield,
uplift

Fig 4: Interaction curves in the normalized NM plane for bearing capacity failure on rigid or deformable soil.
Decomposition of uplifting and soil-yielding mechanisms.

170

onsets whenever the subgrade reaction due to


moment loading at a corner point pm reaches
the initial vertical reaction po . The different
states of the foundation response in terms of a
large-displacement analysis procedure are
presented in Fig 5.

MACROSCOPIC FOUNDATION MODELLING


Elastic soil

For a vertical load factor << 1 (i.e. for a


lightly loaded foundation or for very stiff soil
conditions) foundation rocking is associated
with large levels of uplift and small soil
deformations. Under such circumstances the
assumption of elastic soil behaviour is
generally a reasonable approximation in
analyzing the rocking response. Nonlinear
behaviour under moment loading is then
attributed primarily to the reduction of the
footing contact area (geometric nonlinearity).
In slender structures geometrical nonlinearity
is even more amplified at large rotations due
to the accompanying lateral movement of the
mass centre ( P effect).
In light of an elastic approach for the soil
medium, concentration of contact pressures
occurs in the vicinity of corner points which
cannot be captured by the conventional
Winkler modelling. The increase in local soil
stiffness with the distance from the footing
midpoint is higher under antisymmetric
loading which corresponds to larger values of
rocking stiffness K m compared to the vertical
stiffness Kv . It is therefore evident that a
more efficient and precise soil model should
be incorporated in the nonlinear analysis of
footings subjected to severe overturning
moments. Neglecting the effects of soil
nonlinearity, such a rigorous macroscopic
modelling of rocking behaviour should be in
agreement with: (a) the classical elastic
medium solutions of the soil-foundation
stiffnesses during full-contact conditions (i.e.
Gazetas, 1991), and (b) the limiting case of an
uplifting foundation on a rigid soil.
We consider a rigid, strip footing of width
B = 2b resting on the surface of a
homogeneous half-space. Initially the footing
is subjected only to a vertical loading N .
Then a gradually increasing horizontal force is
applied at the mass center of the
superstructure (located at height h ) leading to
an overturning moment M = Nh cos about
the foundation center. It is postulated that
tensile forces cannot be undertaken by the
soil-foundation interface. In this way uplift

N
M ( < Muplift )

wbo

2b

(a) Full-contact phase


N
Uplift condition :
pz(-b) = 0

M ( = Muplift )

wbo
= uplift
2b

(b) Incipient uplift


Mass
center

N
M ( > Muplift )

wbo

2b

(c) Uplifting phase

Fig 5: Rocking of a rigid strip footing on elastic soil:


different states of response.

Full-contact phase: For sufficiently low


levels of moment, the confinement due to
vertical loading ensures a full-contact
condition at the interface. At this state, the
strip is rocking about its midpoint and
therefore any vertical displacement w at a
distance x from the center comprises (a) a
vertical component w bo and (b) a rocking
component x sin :

w b ( x ) = w bo + x sin

(10)

Moreover, the contact pressure to the footing


at a distance x can be determined with the

171

superposition of the symmetric (vertical) and


the antisymmetric (moment) components of
loading ( pv and pm respectively):

pz ( x, ) = pv ( x ) + pm ( x, )

where Kv , K m are the (global) soil-foundation


stiffnesses for each loading case. These
stiffnesses may be calculated with closedform solutions of the literature. For example
for a rigid strip on elastic half-space it is
Kv = 0.73 G /(1 ) and K m = G b 2 / 2(1 )
(Gazetas, 1991). Recalling Eqs 13 the
subgrade moduli ratio at a distance x from
the center is:

(11)

It is remarked that in contrast to the


conventional winkler modelling, the sensitivity
of soil response to the loading conditions is
now incorporated by adopting two separate
spring constants for vertical and moment
loading.
The problem of a rigid strip on a
homogeneous isotropic medium can be
analyzed in a variety of ways. These include
Greens function techniques, complex variable
methods and integral transform methods.
Within the assumption of a smooth interface,
closed-form solutions can be derived for
simple loading cases. Sadowsky (1928) and
Muskhelishvilli (1953) developed solutions for
the contact pressures in symmetric or antisymmetric loading:
pv ( x ) =

kv ( x ) K v b 2
=
km ( x ) 2K m

which is independent of the distance x .


The sensitivity of the subgrade modulus to
the distance from the midpoint and the loading
conditions is reflected in the p w curves,
calculated at the center and at the edges of
the foundation with two-dimensional finite
element analysis and plotted in Fig 6. The
vertical load which is initially applied to the
footing induces a uniform settlement w bo . The
horizontal (displacementcontrolled) loading
imposed at the level of the mass center
invokes an additional loading in the right part
of the footing (see curve 3) with stiffer
subgrade modulus as well as unloading of the
left part (see curve 1). At this state of
response, symmetric points of the footing with
respect to the vertical central axis have equal
subgrade moduli (i.e. km1 = k m 3 ).

b 1 x 2 / b2

and

(12)
pm ( x, ) =

2M x

b3 1 x 2 / b 2

Evidently, the contact pressure for each


loading case theoretically approaches infinity
at the edges of the footing. In reality the
maximum value of the contact pressure is
bounded due to the finite soil strength and
redistribution of stresses. The subgrade
modulus for vertical and moment loading at a
distance x from the foundation mid-point is
derived after dividing Eqs 12 by w bo and
x sin respectively:

kv ( x ) =

4
km3*

pz / pvo

(3)

(1), (3)

(1)

kv1*, kv3*
kv2*
0

0.25

km1*

(2)
km2*

0.5

0.75

1.25

1.5

w b / w bo

Fig 6: p w curves at the center (2) and at the


edges of the foundation (1, 3) during a
push-over loading test.

2K m

(2)

b 1 x 2 / b2

(1)
2
1

(13)

km ( x ) =

(3)

Kv

and

(14)

1 x 2 / b2

172

The resultant moment of the contact


pressures with reference to the center can be
computed with integration over the foundation:

geometry of the supporting medium and


induces uplift to initiate at a vertical
displacement w ( b ) larger than zero. This is
evident in the numerical analysis depicted in
Fig 6 where uplift onsets before the
normalized vertical displacement of the
unloading (left) edge w bo / w b becomes zero.
Taking into account that at marginal uplift the
vertical displacement of the footing pivot point
is w ( b ) = w uplift the following expression for

M = M p,v + M p,m
b

k ( x )w
v

bo

x cos d x

(15)

x 2 cos sin d x

the rotation at incipient uplift is derived


according to Fig 6:

where M p,v are M p,m the moments associated


with the symmetric and antisymmetric loading
respectively. Setting the transformation
x = b sin ( d x = b cos d ) , these moment
components yield:
b

Mf ,v =

Nx
1 x 2 / b2

Nb cos

/2

sin uplift =

cos d x
(16a)

uplift

sin d = 0

/ 2

Mf ,m =

K m x 2 sin 2

1 x 2 / b2

K m sin 2

w bo kv (b )
b km (b )

(18)

Substitution of Eq 14 into Eq 18 derives the


following uplifting criterion for the rocking
angle:
Mrigid
Nb
=
2K m
2K m

(19a)

Mrigid
Nb
=
2
2

(19b)

and therefore

and
b

w bo w uplift

Muplift

dx

/2

1 cos 2
d

2
/ 2

According to the conventional model, the


uplifting criterion of Eq 17 is satisfied when
w ( b ) = 0 meaning that at the uplift onset the
footing edge reaches its initial position. On the
contrary, from the proposed model uplifting
occurs before the footing edge returns to its
initial position. For a homogeneous soil profile
the arising restitution ratio is depending upon
the footing geometry and the presence of
shallow bedrock.
Moreover, in terms of the conventional
model the critical angle for incipient uplift is
or
uplift N / Kv b
uplift = sin1 (w bo / b )

(16b)

= K m
Through Eqs 16 it is verified that the
symmetric part of loading does not contribute
to the resultant moment whereas a linear
moment-rotation relationship is established
under antisymmetric loading.
Uplift initiation: For a clockwise rotation,
the footing marginally lifts off the supporting
soil when the contact pressure at the left edge
counter-balances the initial reaction of vertical
loading:

pz ( b ) = pv ( b ) pm ( b ) = 0

= Nb / 3K m and therefore the uplifting moment


Evidently,
the
is Muplift = N b / 3 = M rigid / 3 .
conventional Winkler model underestimates
the moment at incipient uplift by a factor of 1.5
in comparison to the exact two-dimensional
solution.

(17)

In contrast to conventional Winkler-based


approach, in the proposed model moment
loading is associated with stiffer subgrade
moduli. This is a result of the two-dimensional

Uplifting phase: Once uplifting occurs the


foundation area remaining in contact with the

173

ground is gradually decreasing (see Fig 5c).


As a result the rocking response enters a non
linear regime even under purely elastic soil
conditions. During the uplift mode, the
moment of the foundation with respect to its
midpoint is the sum of (a) the moment due to
the contact pressures ( M p,v and M p,m ) and (b)

wb

w b = w b + b (1 ) ln

(20)

pm ( x, ) x cos d x

b 2

N h sin

1 2 2

(24)

2M

pm ( , ) =

1 2 2

By virtue of Eqs 21 and 24 the moment due to


purely vertical loading for the uplifting footing
yields:

(21)

The contact pressures of the effective


footing on the basis of elastic soil behaviour
are calculated with the hypothesis of the
incipient uplift:
The distribution of subgrade reactions
along the footing due to combined vertical and
moment loading once uplifting occurs is the
same with that of a fictitious footing of
width 2 , which under the same combined
loading is being at incipient uplift.
A consequence of this hypothesis is that
the center of the effective footing at each step
is the instantaneous rotation pole of the
footing. This can be expressed by the
following incremental equation:

w b = b (1 )

(23b)

pv ( ) =

We define the distance of a point from


the effective footing (the part of the footing
remaining in contact with the ground)
midpoint. In this case:
x = +b

The latter equation provides the analytical


relationship between the vertical displacement
and the rocking angle of the footing over
elastic soil.
The hypothesis of the incipient uplift
allows for the superposition principle to be
applied for the analytical calculation of the
contact pressures developed along the
effective footing:

pv ( x ) x cos d x +

b 2

(23a)

or

M = M p,v + M p,m + MP
b

b (1 )

uplift

w bo

the moment ensued by the lateral movement


of the superstructure ( P effects). In this
way the foundation moment becomes:

wb =

Mf ,v =

N ( + b )

1 2 / 2

/2

N cos

cos d

( sin + b ) d

(25)

/2

= N cos ( b )
This denotes that the moment Mf ,v is the
product of the vertical reaction resultant times
the distance from the effective footing center
to the footing midpoint. The moment of the
foundation due to purely moment loading is:

Mf ,m =

(22)

K m,eff sin 2 ( + b )

3 1 2 / 2

where is the effective-to-initial footing width


ratio ( / b ) .
Now, the vertical displacement of the
foundation w b can be derived by integration:

K m,eff sin 2

174

/2

2
b
sin + sin
d

/2

K m,eff sin 2

K m,eff

(26)

The
above
equation
verifies
the
expression adopted for the subgrade stiffness
due to moment loading.
Finally, the total overturning moment of
the uplifting foundation yields:
M = K m,eff + N ( b ) cos Nh sin

limiting case that the footing is supported by


rigid soil it yields = 0 for any > 0 and
therefore Eq 29 leads to the well-known
momentrotation relationship:
M = Nb cos Nh sin = NR sin ( c )

(27)

While the analytical formulation of the


foundation moment was based on a strip
footing it can also be extrapolated to any
rectangular spread foundation where the
direction of horizontal loading runs parallel to
a normal axis of the foundation cross-section.

However, the rocking stiffness of the effective


footing is:
K m,eff =

G 2
2
= Km 2
2 (1 )
b

(28)

1 : rocking at incipient uplift

By substituting Eq 28 into Eq 27 we finally get:


M = Km

b2

(31)

2 : rocking with uplift

2b sinuplift = 2 sin

uplift

+ N ( b ) cos Nh sin (29a

/ b uplift /

2b sinuplift

2 sin

or
M = K m 2 + Nb (1 ) cos Nh sin

(a)

(29b
)
N = 500 kN,

Finite element analysis of rocking


response reveals that the width of the
effective footing is inversely proportional to
the rocking angle as shown in Fig 7b. In this
way the effective footing ratio may be
calculated with the following equation:

=
=

= 1

uplift

uplift

(30a)

> uplift

(30b)

2b = 2 m,

h=5m

E = 100 MPa - Analytical


-//-

0.8

FE

= /b

E = 20 MPa - Analytical
-//-

0.6

FE

0.4

0.2

0
0

0.02

0.04

0.06

0.08

0.1

: rad

(b)

A geometrical explanation of the relationship


between the effective width and the rocking
angle is shown in Fig 7a.
Eqs 29 and 30 provide the analytical
expression of the foundation moment as a
non linear function of the rocking rotation.
Denotably the above analytical procedure can
be utilized in the computation of the moment:
(a) in the linear domain where full-contact
conditions are encountered, (b) in the largeamplitude region where the gradually
amplifying P
effects dominate the
response, and (c) at near-overturning
conditions where c = tan1(b / h ) . In the

Fig 7: Comparison of the numerical curves of the


- relationship with the analytical
prediction.

A comparison of the analytical calculation


of the foundation moment and vertical
displacement to the results of the finite
element analysis is shown in Fig 8 for a strip
footing ( 2b = 2 m, h = 5 m, N = 500 kN) on
elastic soil layer over rigid bedrock
( E1 = 20 and E2 = 100 MPa ) . An excellent
agreement between analytical and numerical

175

method is achieved throughout the range of


the rocking rotation.
N = 500 kN,

2b = 2 m,

500

h=5m

500

M : kNm

E = 100 MPa
400

400

300

300

200

200

100

100

E = 20 MPa

0
0

0.05

0.1

0.15

0.2

0.05

: rad
0.2

0.2

0.15

0.2

E = 20 MPa

0.15

0.15

0.1

0.1

0.05

0.05

-0.05

-0.05
0

0.05

0.1

0.15

0.2

0.05

0.1

: rad

: rad

500

500

E = 100 MPa

M : kNm

0.15

0.2

E = 100 MPa

wb : m

0.1

: rad

E = 20 MPa

400

400

300

300

200

200

100

100

0
-0.05

0.05

0.1

0.15

0.2

0
-0.05

wb: m

Analytical curves
FE curves

0.05

0.1

0.15

0.2

wb: m

M ( ) = K m 2 + Nb (1 ) cos Nh sin
w b ( ) = w bo + b (1 ) sin + ln (1 )

Rigid soil boundary

uplift / ,
1 ,

> uplift
uplift

Fig 8: Analytical curves of a rigid strip footing on elastic soil (Eqs 22, 29, and 31) and comparison with twodimensional finite element results.

176

During the uplift regime the foundation


moment may be expressed explicitly as a
function of the effective footing ratio by
substituting the rocking angle given by Eq 30b
to Eq 29b:

Nh
M = Muplift 2 2

Km

and the moment at this point:

Nh
Mu = Nb 1

K m

or

(32a)

2u
Mu = Nb 1
tan c

or without P effects

M = Muplift ( 2 )

u
Nb 1 2
c

(35b)

We notice that when u 0 the ultimate


moment approaches the rigid soil limiting
value ( Nb ), while for u c / 2 the moment

(32b)

The latter equation is that presented by


Crmer et al (2002) extracted empirically from
the results of a parametric numerical study.
In the foregoing it was pointed out that
once uplifting initiates the foundation stiffness
enters into a softening fashion which bounds
the overturning moment to an ultimate value.
This upper limit corresponds to the moment
capacity of the foundation. It is well known
that for the extreme case of an infinitely rigid
soil this ultimate moment equals to the vertical
load times the half-width of the footing. In
reality though, soil deformations in the vicinity
of the base edge are inevitable even when
dealing with very stiff soils. Thus, soil
compliance shifts the axis of the resultant
vertical reaction towards the base centre,
reducing the moment capacity of the
foundation. For elastic soil conditions, the
moment Mu is the local maximum of the
function M = M ( ) as defined in Eqs 29.
Hence it can be calculated by the
condition d M / d = 0 . For small values of the
rocking angles (compared to the critical angle
c ) it is sin and cos 1 . Therefore, the
rotation at which the ultimate moment of the
foundation occurs can be computed with
derivation of Eq 29 with respect to :
dM
N 2b2
=
N h
d
4K m 2

(35a)

Mu tends to zero. This means that the locus

of points (u , Mu ) tracks onto the median of


the angle defined by the ordinate and the
rigid M-theta softening line. That conclusion
is validated through the results of finite
element analysis as depicted in Fig 9. The
exact location of an ultimate point (u , Mu ) at
this locus depends on the vertical load, the
height of the mass-point and the soilfoundation rocking stiffness.
600

E = 100 MPa

500

E = 20 MPa
M : kNm

400

E = 5 MPa
300
200
100
0
0

0.05

0.1

0.15

0.2

: rad

Fig 9: Monotonic M curves for strip footing on


elastic soil (5, 20, and 100 MPa) calculated
with finite element analysis and the arising
locus of the critical equilibrium points.
( N = 500 kN, 2b = 2 m, h = 5 m ).

(33)
Inelastic soil

Therefore the angle that satisfies the equality


d M / d = 0 is:

u =

Nb 2
Nh
= tan c
4K m h
4K m

The simplification of elastic soil behaviour


allows for a tractable analytical calculation of
the uplifting response. Excluding some cases
of lightly loaded foundations ( < 0.2 0.3 ) , in
reality yielding zones of the supporting soil

(34)

177

emanating from the area underneath the


foundation edges are rather inevitable. This
may lead to a visible non linear fashion of
the momentrotation relationship even during
full-contact conditions. Moreover, once uplift
initiates,
soil
material
nonlinearity
counterbalances the uplifting displacements
and results into substantially non-linear
foundation behaviour.
We consider again a rigid, strip footing of
width B = 2b resting on the surface of a
homogeneous half-space. Initially the footing
is subjected only to a vertical loading N . The
vertical
displacement
of
the
footing
(settlement) w bo is now a non linear function
of the applied load. Two analytical curves are
most suitable to describe the backbone N w
curve according to the exponential and the
hyperbolic law (Eqs 36a and b respectively):

1200
1000

N: kN

800
600
400
200
0
0

0.08

0.1

0.12

0.1

0.12

1200

800
600
400

(36b)

200
0
0

where = K v / Nu . In Fig 10 the analytical


N w curves are plotted in comparison with
those calculated from finite element analysis
of a strip footing ( 2b = 2 m, h = 5 m ) on
homogeneous
soil
layer
over
rigid
An
bedrock ( E = 100 MPa, su = 100 kPa ) .
excellent agreement is achieved for both
models with the numerical results. According
to the afore-discussed models, the settlement
at the end of this loading phase may be
calculated as follows:
=

0.06

HYPERBOLIC
MODEL

(36a)

ln (1 N / Nu )

0.04

1000

1 e w b

wb
N=
1
1
+
w
Kv Nu b

wb =

0.02

wb: m

N: kN

N = Kv

EXPONENTIAL
MODEL

ln (1 )

Kv / Nu
N
1
1
wb = u
=
Kv 1 Nu / N
(1 1/ )

0.02

0.04

0.06

0.08

wb: m

Fig 10: Comparison of the exponential and the


hyperbolic model with the finite element
analysis in the calculation of the N w
curve during monotonic vertical loading.

At the second loading stage, a gradually


increasing horizontal force is applied at the
level of the superstructure mass (located at
height h ) leading to an overturning moment
M = Nh cos about the foundation center. As
in the elastic case, uplift onsets whenever the
subgrade reaction due to moment loading at a
corner point pm reaches the initial vertical
reaction po .

(37a)
(37b)

Full-contact phase: For sufficiently low


levels of moment, the confinement due to
vertical loading ensures a full-contact
condition at the interface. In contrast to the
elastic case, the strip is rocking about a point
(pole of rotation) which is not fixed at the
footing midpoint but shifts towards its
unloading edge. This leftward (for a clockwise

The analytical calculation of the settlement


with the exponential model has been
successfully evaluated by Nova and
Montrasio (1991) through experimental N w
curves on sand.

178

rotation) movement of the instantaneous pole


is attributed to the plastification of the
supporting soil underneath the loading edge
of the footing. The larger the structural weight
is, the more rapidly the pole moves towards
the unloading edge. In the limiting case of
1 the footing tends immediately to rotate
about its unloading edge even under a very
small overturning moment. In this case the
vertical displacement (settlement) becomes:
w b = b sin

k ( b)
v
w vo

pu 1 e pu
= pu

where kv and km are the elastic stiffnesses


as determined above. Eq 40 may reduce to
the following expression of the stiffness ratio:
w vo w uplift
kv ( b )
=
k m ( b )
w vo

(38)

On the other hand, for elastic soil


behaviour ( = 0 ) the displacement w b prior
to uplift initiation is zero.
The M curve at the full-contact phase
may be calculated with the exponential or the
hyperbolic model as follows (Eqs 39a and 39b
respectively):
M = Km

M=

400
300

(39a)

200
100

1
1
+

K m Mu

(41)

EXPONENTIAL
MODEL

M: kNm

1 e

k ( b)
m

(w bo wuplift )
p
1 e u
(40

0
0

(39b)

0.01

0.02

0.03

0.04

0.03

0.04

: rad

where = K m / Mu and Mu is the ultimate


moment of the foundation if uplift is prevented.
The afore-mentioned analytical models
are evaluated in the calculation of the M
relationship through finite element analysis for
a strip footing ( N = 500 kN, 2b = 2 m, h = 5 m )
on homogeneous soil layer over rigid bedrock
( E = 100 MPa, su = 100 kPa ) as plotted in Fig
11. Both models seem to capture the basic
features of the backbone curve. Nevertheless,
a slightly closer fit is achieved with the
exponential model throughout the loading
sequence.

HYPERBOLIC
MODEL

400

M: kNm

300
200
100
0
0

0.01

0.02

: rad

Fig 11: Comparison of the exponential and the


hyperbolic model with the finite element
analysis in the calculation of the M
curve during monotonic horizontal loading
at the level of the superstructure mass
center.

Uplift initiation: As in the elastic case,


the footing marginally lifts off the supporting
soil when the uplifting criterion described with
Eq 17 is satisfied. By applying the exponential
law to the p w curve of the unloading edge,
the following uplifting criterion is obtained:

From Eq 41 it is derived that the critical


rocking angle at marginal uplift is the same
obtained with the elastic approach:

179

uplift

Nb
2K m

The components of the moment


associated with the eccentricity of the
resultant reaction N ( M p,v ) and the P

(42)

effects ( MP ) are the same with those of the


elastic case.
For the analytical calculation of the
moment component associated with the
antisymmetric part of the external loading
( M p,m ) calculation of soil reactions due to that

The uplifting moment is then derived by


substituting Eq 42 to Eq 39a:
Muplift

Nb

2 Mu
= Mu 1 e

(43)

Evidently, the uplifting moment is not


anymore a linear function of the vertical load
N but exhibits a softening behaviour as N
increases. As the failure curve in the N M
plane is described with a parabola, it is
expected that a threshold value of N exists
where the uplift and failure curves intersect
and beyond which no uplifting occurs.

loading and integration along the effective


footing is required. For inelastic soil conditions
however, this method leads to complex
integral expressions of the resultant moment
even when simple analytical p w curves are
employed (for example the exponential law).
More than that, the shift of the rotation pole
towards the edge of the footing, further
complicates the decoupling of the developed
soil
reactions
into
symmetric
and
antisymmetric components. An alternative
approximate method to calculate the moment
component M p,m may arise through the

Uplifting phase: Once uplifting occurs,


the foundation area remaining in contact with
the ground is gradually decreasing. However
the magnitude of uplifting (i.e. expressed with
the effective footing ratio = / b ) is limited
with the increase in the structural weight
(expressed with the load factor ).
The vertical displacement w b in the
limiting cases of = 0 and = 1 is given by
Eqs 23b and 38 respectively. From these two
states of response, linear interpolation may
provide the vertical displacement w b of a load
factor as follows:
w b w b + b (1 ) ln

extrapolation of the linear M correlation


extracted for elastic soil conditions.
This linear trend of the M correlation
is confirmed by the numerical results plotted
in Fig 12b.
In addition, the effective footing ratio
must satisfy: (a) the limiting case of elastic soil
( = 0 ) and (b) the limit state condition = .
The exponential law of the effective footing
ratio with respect to the rocking angle
which satisfies both limitations is given by the
following equation:

(44a
)

or
w b w b + b (1 ) ln

(44b
)

(46)

Numerical validation of the latter equation


through nonlinear finite element analysis of
rocking response is presented in Fig 12c.
Eventually, the moment of the foundation
for inelastic soil conditions becomes:

During the uplift mode, the moment of the


foundation with respect to its midpoint is the
sum of (a) the moment due to the contact
pressures ( M p,v and M p,m ) and (b) the


M = Km
+ Nb (1 ) cos Nh sin (47
1
)

moment ensued by the lateral movement of


the superstructure ( P effects):
M = M p,v + M p,m + MP

uplift (1 ) +

(45)

180

Eqs 46 and 47 comprise the analytical


expression of the foundation moment as a
non linear function of the rocking rotation for
the general case of inelastic soil.

(50b)

Mu = Nb (1 )(1 )

(51)

in dimensionless form:
Mu

=
(1 )(1 )
2
Nu B
n
= (1 ) (1 n )
m
2

(52a)
(52b)

Eqs 52 provide the analytical expression


of the failure curve in the nondimensional

plane when P
effects are
n m
considered. Remark that for = 0 (i.e. no
P effects are considered) the failure curve
reduces to that calculated with the
conventional Winkler model (Eq 8).

CONCLUSIONS

Non linear features of the rocking


response of tall structures founded on shallow
foundations are investigated. Principally,
based on conventional Winkler modelling,
interaction curves in the nondimensional
N M plane were calculated for failure
(overturn) as well as incipient uplift and soil
yield. A perfectly symmetric response is
achieved about the vertical axis at = 0.5 . At
this point the optimum foundation behaviour in
terms of the moment ultimate capacity is
obtained.
To
highlight
the
induced
nonlinearities in the large-displacement
domain, a macroscopic modelling of the soil
foundation system was developed. In this
respect
analytical
equations
for
the

(48)

The rocking angle u at which the ultimate


moment is attained is therefore:
Nh
4K m

2u
u
Mu = Nb 1
Nb 1 2
c
tan c

After some algebraic manipulations, the


ultimate moment may be expressed as a
function of the nonlinear parameters
= Nh / K m
and , which represent
respectively geometrical and soil material
nonlinearities.

Interaction in the N M plane: As


shown in the foregoing discussion, the M
backbone curve follows a softening fashion
due to geometric and soil material
nonlinearities. As a result the moment
developed by the foundation is bounded by an
ultimate value Mu . In the optimum case of a
rigid supporting soil the foundation undertakes
the maximum possible moment which
corresponds to the vertical load N times the
half-width of the footing b . In common
geotechnical applications however soil
compliance and plastification of the supporting
soil result in substantial reduction of the
ultimate moment. Moreover, further decrease
in Mu is expected in tall structures due to
P effects. In the general case the ultimate
moment of the foundation may be derived
analytically as the local maximum of Eq 47.
To this extent derivation of Eq 47 yields:

u = tan c

(50a)

or

A comparison of the analytical calculation


of the foundation moment and vertical
displacement to the results of the finite
element analysis is shown in Fig 13 for a strip
footing ( 2b = 2 m, h = 5 m, N = 500 kN) on
inelastic soil layer over rigid bedrock
( E = 100 MPa and su = 100 kPa ) .
A close agreement of the analytical results to
those obtained from numerical analysis is
attained throughout the range of the rocking
rotation.

dM
N 2b2
=
N h =0
d
4K m 2

Nh
Mu = Nb 1

K m

(49)

Eventually, the ultimate moment is:

181

to uplift, (b) M correlation, and (c) the effective


width 2 with respect to the rocking angle and
comparison with the analytical prediction.

overturning moment and the vertical


displacement were extracted as a function of
the rocking angle (a) for elastic soil, and (b)
for inelastic soil. From the latter case,
analytical equation of the failure curve in the
nondimensional N M plane was obtained,
incorporating both geometric and material
nonlinearities.

E = 100 MPa
400

M : kNm

N = 500 kN,

500

E = 100 MPa,

2b = 2 m,

0.5

h = 5 m,

300
200
100

120

0
0

gap force: kN

100

0.05

0.1

0.15

0.2

: rad

80

xp
0.2

60

E = 100 MPa

40

0.15

20
-1

-0.5

0.5

wb : m

0
1

distance from axis: m

0.1
0.05

(a)
0

400

-0.05
0

M: kNm

300

0.02

0.04

0.06

0.08

0.1

: rad

200

500

Muplift

E = 100 MPa

100

400

0.5

0.6

0.7

0.8

0.9

M : kNm

0
1

300
200

(b)
100

2.00

0
-0.016

-0.014

2: m

-0.012

-0.01

-0.008

wb: m

1.50


M = Km
+ Nb(1 )cos Nh sin
1

1.00
Analytical curves
FE curves

0.50

Rigid soil boundary

0.00
0

0.005

0.01

0.015

w b w b + b (1 ) ln
uplift + ( uplift )

0.02

: rad

1,

> uplift

uplift

(c)

Fig 13: Analytical curves of a rigid strip footing on


inelastic soil and comparison with twodimensional finite element results.

Fig 12: Nonlinear rocking response of a strip


footing on inelastic soil, calculated with twodimensional finite element analysis: (a) distribution
of soil reactions in characteristic increments prior

182

Crmer, C., Pecker, A. and Davenne, L. [2002].


Modeling of nonlinear dynamic behaviour of a
shallow strip foundation with macro-element,
J. Earth. Engng, 6(2).

ACKNOWLEDGEMENTS

The major part of the present study was


funded by the Greek General Secretariat for
Research and Technology through the
research project X-SOILS. Also part of this
work was conducted in terms of the EU
research project QUAKER, funded through
the EU Fifth Framework Programme:
Environment, Energy, and Sustainable
Development, Research and Technological
Development Activity of Generic Nature: the
Fight against Natural and Technological
Hazards (contract number: EVG1-CT-200200064).

FEMA 356 [2000]. Prestandard and commentary


for the seismic rehabilitation of buildings,
FEMA, Washington, D.C.
Gajan S, Kutter B, Phalen J, Hutchinson T and
Martin G. [2005]. Centrifuge modelling of loaddeformation behaviour of rocking shallow
foundations, Soil Dyn. & Earth. Engrg 25, 773783.
Gazetas, G. [1991]. Foundations vibrations,
Foundation Engineering Handbook ch 15 (H.Y. Fang, van Nostrand Reinhold, NY).
Gazetas G., Apostolou M. and Anastasopoulos I.
[2003], Seismic Uplifting of Foundations on
Soft Soil, with Examples from Adapazari (Izmit
1999, Earthquake), BGA International Conf. on
Foundations Innovations, Observations,
Design & Practice in the University of Dundee,
Scotland, September 25, pp. 37-50.

REFERENCES

Allotey, N. and El Naggar, M. H. [2003]. Analytical


moment-rotation curves for rigid foundations
based on a Winkler model, Soil Dyn. & Earth.
Engrg 23, 367-381.

Gazetas, G. and Apostolou, M. [2004]. Nonlinear


soilfoundation interaction: Uplifting and soil
yielding, Proc. 3rd Joint US-Japan Workshop
on SSI, Menlo Park CA.

Apostolou M., Gazetas G., Makris N.,


Anastasopoulos I. [2003], Rocking of
Foundations under Strong Seismic Excitation.
Proc. Fib International Symposium on Concrete
Structures in Seismic Regions, Athens, May
2003.

Gazetas G., Apostolou M., and Anastasopoulos I.


[2004], Seismic Bearing Capacity and Uplifting
Foundations : Adapazari 1999, Proc. 5th Int.
Conf. on Case Histories in Geotechnical
Engineering, New York, April 13-17.

Apostolou, M., Thorel, L., Gazetas, G., Garnier, J.


and Rault, G. [2007]. Physical and numerical
modelling of soil-footing-structure under
lateral cyclic loading, Proc. 4th ICEGE,
Thessaloniki, Greece.

Gazetas, G. and Apostolou, M. [2006]. Numerical


modelling of shallow foundations, Quaker
program, Final report on Topic B2.

ATC 40 [1996]. The seismic evaluation and retrofit


of concrete buildings, Applied Technology
Council. Vols. 1 and II, Palo Alto, CA.

Gerolymos, N., Apostolou, M., and Gazetas, G.


[2007], Neural network analysis of overturning
response under nearfault type excitation,
Earthquake Engineering and Engineering
Vibration, Vol. 4, No. 2, pp. 213-228.

Bartlett, P. [1979]. Foundation rocking on a clay


soil, M.E. Thesis, Report No. 154, School of
Eng., Univ. of Auckland, 144 pp.

Hibbitt,
Karlsson
and
Sorensen
[2002].
Abaqus/Standard Users Manual, Version 6.3,
Hibbitt Providence, Rhode Island.

Bransby, F. and Randolph, M. [1997]. Shallow


foundations subjected to combined loadings,
Proc. ISOPE.

Houlsby G.T. and Puzrin A.M. [1999]. The bearing


capacity of strip footing on clay under
combined loading, Proc. Royal Society, 455A:
893-916.

Butterfield, R. and Ticof, J. [1979]. The use of


physical models in design, Discussion. Proc.
7th Eur. Conference Soil Mechanics, Brighton,
4, 259-261.

Martin, G. and Lam, I. [2000]. Earthquake


resistant design of foundations Retrofit of
existing
foundations,
Proc.
GeoEng
Conference, Melbourne.

Butterfield, R. and Gottardi, G. [1994]. A complete


three-dimensional failure envelope for shallow
footings on sand, Gotechnique 44, No. 1,
181-184.

183

Nova, R. and Montrasio, L. [1991]. Settlements of


shallow foundations on sand, Gotechnique
41 (2), 243-256.
Pecker, A. [1998] Capacity design principles for
shallow foundations in seismic areas, Keynote
lecture, Proc. XIth Europ. Conf. on Earthquake
Engrg. Eds. Bisch, Labb, Pecker Balkema.
Priestley, M. J. N., Evison, R. J. and Carr, A. J.
[1978]. Seismic response of structures free to
rock on their foundations, Bull. New Zealand
Nat. Soc. Earth. Engrg. 11(3), 141-150.
Sadowsky,
M.
[1928].
Zweidimensionale
probleme der Elastizittstheorie, Zeit. Angew.
Math. Mech. 8(2), 107-121.
Siddharthan, R., Ara, S., Norris, G. [1992]. Simple
rigid plastic model for seismic tilting of rigid
walls, J. Struct. Engng, ASCE 118(2), 469487.
Somerville, P. G. [1998]. The amplitude and
duration effects of rupture directivity in near
fault ground motions, Geot. Earth. Engrg. Soil
Dyn. III ASCE, eds. P. Dakoulas, M. K. Yegian
and R. D. Holtz.
Quaker project website:
www.dundee.ac.uk/civileng/quaker/index.htm.

184

Natural Period and Effective Damping


of Simple Structures on Footings and Piles
A. Maravas, G. Mylonakis and D. L. Karabalis
University of Patras, Greece

Abstract
Novel analytical solutions are presented for single degree-of-freedom (SDOF) oscillators founded on footings
and piles on compliant ground. First, exact formulas for the fundamental natural period of the above
structures, encompassing the frequency dependence of the various impedance terms, are derived. Second,
closed-form solutions for the corresponding damping coefficients are derived. It is shown that the common
approximation of neglecting higher-order terms involving products of damping coefficients is unnecessary and
potentially inaccurate for highly-damped soil-structure systems. Third, the influence of foundation mass on the
period and damping of the system is incorporated. To address the issue of coupled swaying-rocking
oscillations at the pile head, the reference system is translated to the depth below the pile head where the
resultant soil reaction to the pile is applied, to ensure a diagonal foundation impedance matrix. Fourth, the
amounts of radiation damping generated from a single pile and a surface footing are compared. To this end, a
new concept of statically and geometrically equivalent SSI systems is introduced. It is shown that a structure
founded on a pile may generate twice the amount of radiation damping produced by a similar structure on a
spread footing. Results are provided in ready-to-use graphs and charts that elucidate the salient features of
the problem and can be directly implemented in design. The paper complements and extends the seminal
studies in the subject by Parmelee, Veletsos, Bielak and co-workers.

(1985), (2) to present a novel solution for


determining the fundamental natural period and
effective damping of a simple oscillator
supported on a surface foundation, (3) to
extend the solution to encompass pile
foundations, (4) to present results for typical
structural systems on compliant ground, and (5)
to compare radiation damping generated from a
surface foundation and a pile foundation.

INTRODUCTION

Knowledge in the subject of dynamic SoilStructure Interaction (SSI) has been derived
mainly from studies of structures on mat
foundations, during the last forty years. The
seismic response of structures on pile
foundations has received considerably less
research attention. More importantly, the results
of these efforts have not yet lead to established
design methods and/or code provisions, such
as the simple methods developed for structures
on surface foundations (NEHRP-03, EC-8).
Therefore much is yet to be learned on the
subject before a comprehensive understanding
is developed on the role of basic problem
parameters on the seismic response of pilesupported systems.
The goals of this article are: (1) to review
available methods on the subject with emphasis
on the design-oriented solutions by Veletsos
and co-workers (1974, 1975, 1977) and Wolf

STRUCTURE ON SURFACE FOOTING

The classical approach for elastodynamic


analysis of soil-structure interaction aims at
replacing the actual structure by an equivalent
simple oscillator supported on a set of
frequency-dependent springs and dashpots
accounting for the stiffness and damping of the
soil medium. This model has been adopted by
several researchers, including Parmelee (1967),
Veletsos et al (1974, 1975, 1977), Jennings &
Bielak (1973), Wolf (1985) and, more recently,

185

Aviles et al (1996, 1998). A brief overview of


available methods leading to closed-form
solutions is presented below. Based on these
procedures,
a
novel,
accurate
and
straightforward scheme for analyzing the
problem is presented.

( ) =

K x = ax K ,

The system studied is shown in Figure 1. It


involves a simple oscillator on flexible base
representing a single storey structure, or a multi
storey structure after a pertinent reduction of its
degrees-of-freedom (e.g., considering that the
mass is concentrated at the point where the
resultant inertial force acts).
The structure is described by its stiffness k,
mass m, height h, and damping ratio , which
may be either viscous or linearly hysteretic. The
foundation consists of a rigid surface circular
footing of radius r resting on a homogeneous,
linearly elastic, isotropic halfspace described by
a shear modulus Gs, mass density s, Poissons
ratio s, and hysteretic damping ratio s.
Foundation stiffness is modeled by frequencydependent springs Kx and K representing
stiffness in translational and rocking oscillations,
respectively. Following Veletsos & Nair (1974),
to ensure uniform units in all stiffness terms, K
is expressed by a translational vertical spring
acting at distance r from the center of the
footing. Damping is modeled by a pair of
dashpots Cx, C, attached in parallel to the
springs, representing energy loss due to
hysteretic action and wave radiation in the soil
medium. In the present formulation, the
influence of foundation embedment on the
response is neglected.
The dynamic impedance K * ( ) along any
degree of freedom of the system is defined
according to the formula

K ( ) = K + i C = K (1 + 2 i )

(2)

For the model in Figure 1a, the foundation


springs and dashpots can be expressed by the
formulas proposed by Veletsos & Meek (1974):

Classical solutions

Im( K * ) C
=
2 Re( K * ) 2 K

Cx = x

Kr
,
Vs

K = a K
C =

(3)

Kr
Vs

(4)

where Vs denotes the propagation velocity of


distortional waves in the halfspace and K is the
static horizontal stiffness of the foundation
defined by

K=

8
Gs r
2 s

(5)

x, , x and are dimensionless factors that


depend on Poissons ratio for the halfspace
material, and the dimensionless frequency
a0 =

(6)

Vs

Under seismic excitation, the system


deflects as shown in Figure 2. The translation of
the mass relative to ground is composed of
three parts: (1) horizontal translation due to
swaying motion of footing ux, (2) horizontal
translation due to rocking motion of footing, u
and (3) horizontal deflection of column, uc.
Based on these definitions, the impedance of
the system is defined as:

j*
K

(1)

P
i 1 + 2ii
=K
u x + u h + uc

(7)

i and i denote the apparent stiffness


where K
and damping coefficients at elevation h.
The response of the soil-structure system
depends on the mechanical properties of the
foundation, the soil, the superstructure and the
characteristics of the excitation. These are
summarized in the following dimensionless
parameters (Veletsos et al 1974, 1975, 1977):
(i) The wave parameter

in which K is the real part of the impedance,


C is the corresponding imaginary part,
is the cyclic excitation frequency, and i
(= 1 ) the imaginary unity. is an energy
loss parameter, analogous (yet not
identical) to the viscous damping coefficient
of a simple oscillator.

186

Vs
fc h

Solution by Wolf (1985)

(8)

The system considered by Wolf is identical


to that shown in Figures 1 and 2. The main
difference with the Veletsos approach is that
frequency-independent moduli defined by the
values ax = 1, x = 0.575 , a = 0.15, = 0.15 , are
adopted for the foundation, and that the
response of the system is evaluated by directly
solving a set of three simultaneous governing
equations of the system.
The properties of the replacement oscillator
in this solution are given by

where fc = k / m / 2 denotes the natural


frequency of the fixed base structure.
(ii) The relative mass density for the structure
and the soil

m
s hr 2

(9)

(iii) The damping ratio of the structure for fixed


base conditions.
(iv) The slenderness ratio (h/r).
(v) The Poissons ratio s of the soil.
(vi) The hysteretic damping ratio s of the soil.

 =

The aim of these solutions is to connect the


properties of the soil-structure system ( Ti , i )
with the properties of the fixed base structure
(, ), so that the influence of soil-structure
interaction on the dynamic behavior of the
structure can be elucidated. This connection is
expressed by the following pair of equations
(Veletsos 1977):
2

( )

( )

i
i = j0 + T T

cyclic natural frequencies of the system under


rocking oscillations of the base (superstructure
assumed rigid), swaying oscillations of the base,
and
oscillations
of
the
superstructure
(foundation assumed rigid), respectively. Note
the simpler form of Eqn (14) as compared to
Eqn (12).
Notwithstanding the theoretical significance
and practical appeal of the above methods,
they both can be criticized on the following
important aspects:

(11)

(a) Both methods neglect products of damping


ratios (i x j) as negligible higher order
terms. This approximation is questionable
for highly-damped SSI systems.
(b) The effective damping in the Veletsos
approach arises from an approximate
procedure leading to an expression
containing imaginary terms (Eqn 12). This
limits significantly its suitability for practical
applications.
(c) Structural damping in the Veletsos solution
is strictly of viscous nature.
(d) Frequency dependence of foundation
springs and dashpots in the Wolf approach
is neglected.

radiation damping of the footing. The latter is


given by (Veletsos and Nair, 1975)

( )

4 Ti
2 3 T

(14)

x = K x / m , c = k / m define the uncoupled

where represents the damping of the structure


(assumed to be of viscous nature) and j the

j0 =

(13)

In the above equations, = K r 2 / mh 2 ,

1 a h
1+3 s x (10)
2s a r

k kh 2
+

K x K


i 2
i 2
i 2
i 2

+
+


x
s

c
c
x

Solution by Veletsos and co-workers (1974, 1975,


1977)

k Kh
2
Ti =T 1+ 1+ x =T 1+ s
Kx K
2 ax 2 h
r

i = 2 1+

(2 s )x r 2
3(1 s )

(12)
+
ax (ax + ia0 x ) h a (a + ia0 )

The method is based on setting the resonant


period and peak pseudo-acceleration of the
actual infinite-degree-of-freedom elastodynamic
system equal to that of an equivalent simple
oscillator. More discussion is given below.

187

(e) Structural damping in the Wolf solution is


strictly of hysteretic nature.
(f)
Both solutions employ rather complex
procedures involving either equivalence
of responses of different dynamic systems
(Veletsos), or solutions, of simultaneous
linear equations (Wolf).
(g) In both solutions, foundation mass and
rotational inertia are neglected.

2
2

i2
1 + 4 i
1 + 4 i
i 2 = 1 + 4

+
+
2
2
2
2
2
2
x 1 + 4 x
c (1 + 4 )
1 + 4

(18)

The above equations are exact solutions to the


problem, in the sense that no approximations
apart from those typically involved in
numerically
evaluating
the
foundation
impedances are involved. Given the frequencydependnet nature of foundation impedances, an
iterative procedure is generally required to
implement these formulas. Note that omitting
the products i2, the above solutions duly reduce
to those in Eqns (14) and (13), respectively.

Proposed exact procedure

In this section a simple exact solution to the


problem shown in Figures 1 and 2 is presented.
The solution contains no approximations in the
derivation of the fundamental natural period and
effective damping of the system. Furthermore,
the
exact
frequency-varying
foundation
impedances may be employed.
Mention has already been made that the
total horizontal deflection of the system can be
decomposed as sum of the three modular
displacements shown in Fig 2, i.e.,

Influence of foundation mass

he influence of foundation mass and inertia


(mf, If) on dynamic behavior of the system
presented in Figures 1 and 2 can be
incorporated by using modified frequencydependent springs Kx and K according to:

(15)

K x = Kx mf 2

(19)

This implies that the associated compliances


can be viewed as complex springs attached in
parallel and, thereby, the dynamic impedance
of the system can be expressed through the
summation rule

K = K f 2

(20)

ut = uc + u x + u

where Kx and K are given by Eqn (3). This


modification does not alter the form of the
solution in Eqs (17) and (18), yet introduces an
additional dimensionless parameter, , which
expresses the ratio of footing mass to
superstructure mass
m
= f
(21)
m

1
1
1 h
1
= * + * + *
*
i
Kx
K r k
K

(16)

in which the associated impedances are


complex valued (to account for phase
differences in the response of the springs) and
frequency
dependent.
Substituting
each
complex impedance term in Eq. (16) by its
representation according to Eq. (1) yields the
exact damping and natural frequency of the
system as (Maravas 2006)

i =

x 2 1 + 4 x 2

x 1 + 4 x
2

2 1 + 4 2

1 + 4
2

which varies typically between 0 and 1. More


discussion is given below.
Parametric analysis and comparisons with
classical methods

Comparative graphs for the variation of the


i / ) and the effective
ratio Ti / T (inverse of
c
damping of the system i , versus the

c 2 (1 + 4 2 )
1
c (1 + 4 2 )

(17)

slenderness ratio h/r, are presented in Figure 3.


Using the proposed exact procedure, the
influence of the relative mass ratio and the
material hysteretic damping ratio s on the
period and effective damping of the soilstructure system is presented in Figure 4. For

188

the parametric analysis, values of factors x, ,


x and correspond to a Poissons ratio of
0.45.
It is evident from Figures 3(a,b) that the
results from the method of Veletsos are in
relative agreement with those obtained by the
proposed exact procedure. Also in Figures
3(c,d), we observe that results from the method
of Wolf are quite different from those of the
proposed procedure. This is mainly due to the
assumption of frequency-independent springs
and dashpots adopted by Wolf.
The above observation is justified by the
fact that for low values of ratio 1/ (i.e., low
values of c) the results of the two methods are
virtually identical. It is also apparent that the
results obtained by Wolf loose accuracy with
decreasing h/r.
Figures 4(a,b) show that the variation in
relative mass ratio affects significantly the
period and damping of the soil-structure system.
More specifically, increasing leads to more
flexible systems and higher values of damping
ratio. The hysteretic damping ratio of soil s,
does not affect much the system period,
especially for tall structures (h/r=5), as shown in
Figure 4(c). On the other hand, it affects
considerably the effective damping of the
system, as shown in Figure 4(d).
Figures 5(a,b) show that the variation of
mass ratio affects the natural period and
damping of soil-structure systems with low
values of h/r. On the contrary, it does not affect
systems with high values of the slenderness
ratio, as shown in Figures 5(c,d).
STRUCTURE ON PILE FOUNDATION

The case of a structure on a single pile


foundation is investigated next. The problem is
treated using a variance of the method for
spread footings presented in the previous
section. The amounts of energy radiated by a
pile- or a footing-supported structure are
compared via by the concept of statically
equivalent SSI systems, introduced in this work.
Soil-pile-structure system and method of
analyses

The system studied is shown in Figure 6. It


consists of the linear elastic SDOF structure
described in previous section, founded on a
single flexible, circular solid pile of Youngs

189

modulus Ep, diameter d, mass per unit length


mp, and length L, which is considered to be
greater than the pile effective length Le.
Accordingly, the pile can be considered
infinitely long. The soil is considered as a
linearly
elastic
homogeneous
isotropic
halfspace, as described above.
Soil-pile system can be represented by
three dynamic impedances K xx* , K rr * , and K xr * ,
corresponding to swaying, rocking, and crossswaying-rocking of the pile head, respectively.
In this study, analytical expressions for the
dynamic impedances are used, as derived by
Novak, (1974) and Mylonakis (1995).
K xx = 4 E p I p 3 , K xr = 2 E p I p 2 , K rr = 2 E p I p
*

k x m p 2 + i cx
=

4E p I p

(22)

1/ 4

(23)

where is a wave number parameter, Ip is the


moment of inertia of the pile cross section, and
kx and cx are the moduli of distributed springs
and dashpots along the pile. The latter are
given by (Gazetas & Dobry 1984):

k x = Es ,

cx = 6a -1/4
op sVs d + 2

s kx

(24)

where aop=2ao (where the footing radius r in Eq.


(6) is replaced by the pile diameter d) and is
the Winkler factor, given as function of pile-soil
stiffness ratio Ep/Es (Dobry et al, 1982)

Ep

Es

= 1.67

0.053

(25)

Under horizontal loading the soil reacts in


the manner shown in Figure 6(b). The resultant
of the distributed reactions is applied at depth e
below the pile head. Since the reference
system is anchored at the pile head, a cross
i xr is
swaying-rocking impedance term K
necessary for modeling the compliance of the
foundation. This term is not compatible with the
analysis of the spread footing presented earlier.
In order to overcome this problem, the
reference system can be translated to a depth
e, where the total soil reaction is applied. In this
manner, the cross impedance K xr * vanishes

and the impedance matrix of the pile becomes


diagonal, as shown in Figure 6(c). This
transformation is approximate, as it implies a
rigid pile between depths z = 0 and z = e.
However, this introduces little error, since e is
usually small (up to a few pile diameters)
compared to the overall pile length. The
transformed impedances K xxe* and K rre* are
given by the expressions
*

K xxe = K xx , K rre = K rr 2 K xr e + K xx e 2 , K xre = 0

Results of parametric analyses

The response of the soil-pile-structure


system depends on the properties of the pile,
the supporting soil, the superstructure and the
excitation. These properties are included in
above dimensionless parameters, as in the
case of the structure supported on surface
footing.
In Figure 7, results obtained with the use of
the proposed exact procedure are presented.
Specifically, Figure 7(a) presents the influence
of the slenderness ratio (h/d) on the natural
period of the interacting system. Evidently, the
behavior is similar to the behavior of the soilfooting-structure system presented earlier. The
same observation holds for the effective
damping, as shown in Figure 7(b).
The influence of pile-soil stiffness ratio Ep/Es,
on the properties of the system is significant, as
shown in Figures 7(c,d). For relatively flexible
piles (low values of Ep/Es), the soil-pile-structure
system is more flexible and dissipates larger
amounts of energy.

(26)

where
*

K
1
e = xr * =
2
K xx

(27)

is the aforementioned eccentricity.


The transformation shown in Figure 6(c)
allows usage of the exact procedure developed
for the analysis of structures on surface
footings. Thus, the natural frequencies i of the
soil-pile-structure system required by Eqs. (17)
and (18) are computed by Maravas (2006) as

K xx
K rr
, =
m
m( h + e) 2

x =

Comparison of two SSI systems

In this section, the concept of statically and


geometrically equivalent interacting systems is
introduced, in an effort to compare the damping
of systems on pile or surface footing
foundations. To achieve this, the lateral
stiffness of the two SSI systems should be
comparable at the elevation of mass m.
Accordingly, the two systems are geometrically
equivalent, i.e., h/d=h/2r, if the ratio of Youngs
moduli for the soil (Maravas 2006):

(28)

where
K xx = ( 4 E p I p )

1/ 4

K rr =

( k m 2 ) 2 + ( c ) 2
p
x
x

3/8

3
cos (29)
4

1/ 8
2
3/ 4
1
2
1
4 E p I p ) ( k x m p 2 ) + (cx ) cos
(

4
4

cx

k x m p 2

= Arc tan

(30)
1/ 4

(f)
Es
1 3
=

( p)
2 16
Es

(31)

= tan
2
4

(2 vs )(1 + vs ) 1 + 4 ( h / d )

1/ 4
1/ 4

16 E p
h

1 + 1 + 2
( p)
d

Es

(33)

In the above equation, symbol f denotes


footing and symbol p pile. Because a pile is
much stiffer than a footing, the supporting soil
of the footing must have a modulus of
(f)
elasticity Es quite larger than the modulus of

The corresponding damping ratios i are given


by the expressions (Maravas 2006)
x = tan ,
2
4

1/ 4

Ep
( p )
Es

(32)

( p)

elasticity of the soil around a pile Es .


Using the proposed exact procedure, the
effective damping of the two SSI systems is
compared for different ratios h/d and Ep/Es, as
shown in Figures 8a and 8b. Furthermore,

in which is defined by Eq. (31).

190

comparative graphs of radiation damping for the


two foundation types are presented in Figures
8c and 8d. Evidently, a structure on a pile may
experience as much as 3 times the amount of
damping generated from the same structure on
a spread footing (Figure 8a). This observation is
justified in view of the tow-dimensional nature of
wave propagation around a pile, which results
to much higher energy dissipation.

REFERENCES
Aviles, J. & Perez-Rocha, L.E.(1996) Evaluation of
interaction effects on the system period and the
system damping due to foundation embedment
and layer depth Soil Dynamics & Earthquake
Engineering. 15(11), 27.
Aviles, J. & Perez-Rocha, L. E (1998). Effects of
foundation embedment during building-soil
interaction, Earthquake Engineering & Structural
Dynamics, 27(12), 1523-1540.

CONCLUSIONS

Dobry, R., Vicente, E., O'Rourke, M. J., & Roesset,


J. M (1982). "Horizontal stiffness and damping of
single piles", J. Geotech. Engng Div., ASCE,
108(3), 439-459.

A novel analytical procedure for determining


the dynamic characteristics of simple structures
founded on surface footings and piles was
presented. Using the proposed methodology,
the influence of common assumptions on the
computation of the mechanical properties of
such systems was elucidated. Results were
provided in ready-to-use graphs and charts that
elucidate the salient features of the problem
and can be directly implemented in design. By
introducing the concept of statically and
geometrically equivalent SSI systems, the
amounts of radiation damping generated from a
single pile and a footing were compared. The
main conclusions of the study are:
(1) The proposed solution is simpler, more
accurate, and more general than the classical
methods by Veletsos, Bielak and co-workers.
(2) The common approximation of
neglecting
higher-order
terms
involving
products of damping coefficients may be
inaccurate for highly-damped SSI systems.
(3)
The
foundation
mass
affects
considerably dynamic SSI response of systems
with low values of slenderness ratio h/r.
(4) The proposed analysis can easily
incorporate
embedded
foundations,
by
translating the reference system to the depth
below the surface where the resultant soil
reaction is applied. This ensures a diagonal
foundation impedance matrix and greatly
simplifies calculations.
(5) A structure founded on a pile may
generate 100% more radiation damping than a
similar structure on a spread footing. The
difference gets more pronounced with highfrequency, squatty structures on stiff soil.

Eurocode EC-8 (1990) Structures in seismic


regions, Part 5: Foundations, retaining structures,
and
geotechnical
aspects,
Brussels:
Commission of the European Communities
Gazetas, G and Dobry, R. (1984) Simple Radiation
Damping Model for Piles and Footings, Journal
of Engineering Mechanics, Vol. 110, No. 6, June,
pp. 937-956
Gazetas G., Anastasopoulos I., Gerolymos N.,
Mylonakis G., & Syngros C. (2005), The
Collapse of the Hanshin Expressway (Fukae)
Bridge, Kobe 1995 : SoilFoundationStructure
Interaction, Reconstruction, Seismic Isolation,
Entwicklungen
in
der
Bodenmechanik,
Bodendynamik und Geotechnik, Festschrift zum
60. Geburstag von Univ.-Professor Dr.-Ing.habil.
Stavros A. Savidis, Frabk Rackwitz, Springer,
pp. 93-120.
Gerolymos, N., (1997), Effective period and damping
of bridge piers on pile groups, Diploma Thesis,
School of Civil Engineering, National Technical
University, Athens (in Greek).
Gerolymos, N., Gazetas, G. & Mylonakis, G.
Fundamental natural period and effective
damping of pile-supported bridge piers, Proc.,
11th European Conference on Earthquake
Engineering, Balkema, Rotterdam, 1998
Jennings, P.C. & Bielak, J.(1973) "Dynamics of
building-soil interaction", Bulletin of the
Seismological Society of America, 63(1) 9-48.
Maravas, A. (2006). Discrete model for dynamic
soil-structure interaction of structures on rigid
surface or pile foundations, MS Thesis,
University of Patras, Greece (in Greek).
Mylonakis, G. (1995) Contributions to the static and
seismic analysis of pile-supported bridge piers,
Ph.D. Dissertation, SUNY-Buffalo.
NEHRP (1997) Recommended Provisions for
Seismic Regulations for New Buildings and other

191

Veletsos, A.S. & Meek, J.W. (1974). Dynamic


behavior of building-foundation systems, Earthq,
Engng & Struct, Dyn,, 3(2), 121-138.

Structures, Building Seismic Safety Council,


Washington, D.C.
Novak, M. (1974) Dynamic stiffness and damping of
piles, Canadian Geotech. Journal, 11, 574-591.

Veletsos, A.S. & Nair, V.V. (1975) Seismic


interaction
of
structures
on
hysteretic
foundations, Journal of Structural Engineering,
ASCE, 101(1), 109-129.

Parmelee, R. (1967). Building-foundation interaction


effects, Journal of Engineering Mechanics
Division, ASCE, 93(EM2), 131-152.

Veletsos, A. S. (1977) Dynamics of StructureFoundation Systems, in: Hall, W. J. (ed.),


Structural & Geotechnical Mech., Prentice-Hall.

Sextos, A, Kappos. A, Pitilakis, K. Inelastic dynamic


analysis of RC bridges accounting for spatial
variability of ground motion, site effects and soilstructure interaction phenomena. Part 2:
Parametric study, Earthquake Engineering and
Structural Dynamics 32 (4), pp. 629-652

Wolf, J. P. (1985). Dynamic


Interaction, Prentice Hall.

Stewart, JP, Fenves, GL, Seed, RB. (1999).


Seismic Soil-Structure Interaction in Buildings I:
Analytical Methods, Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, 125(1).

192

Soil-Structure

FIGURES

u
m

C
r

(b)

(a)

Fig 1: (a) Structure idealized by a stick model, (b) Reduced single degree-of-freedom model

ux

Cx

Kx

Fig 2:. Deflection diagram for soil-structure system

193

uc

h/r=5

2.2

h/r=3

0.4

0.2

2.0

0.1

1.8

1.6

h/r=1

0.3

h/r=1

Exact procedure
Veletsos (1977)

h/r=2

0.05
0.04

0.03

1.4

0.02

1.2

0.01

Exact procedure
- - - - - Veletsos (1977)
h/r=5

1/

1.0
0.1

(a)

0.2

0.3

0.4

0.5

= 0.05, s = 0.45, = 0.15, s = 0

1.0

0.005
0.02

0.6

(b)

h/r=1

0.05

0.10

0.25

0.50

1/

0.75 1.00

= 0.05, s = 0.45, = 0.15, s = 0


0.5

h/r=0.33

0.8

0.4

h/r=5

c 0.6

h/r=2

h/r=1

0.3

0.4

0.2

Exact procedure
Wolf (1985)

Exact procedure
Wolf (1985)

h/r=0.33

0.2

0.0
0.1

(c)

h/r=2

0.1

h/r=5

1/

0.0

1/

0.1

(d) = 0.02, s = 0.45, = 0.15, s = 0.05


Fig 3:. Comparison of proposed exact solution with those of Veletsos (1977) and Wolf (1985)

= 0.02, s = 0.45, = 0.15, s = 0.05

3.5

0.7

3.0

2.5

=0
0.10
0.20
0.30
0.60
1

2.0

1.5

1.0

0.5
0.2

(a)

0.4

0.6

0.8

1.0

1/

= 0.02, s = 0.45, h/r = 1, s = 0.05

0.01
0.1

(b)

3.0
0.1

1/

0.5

h/r=1

= 0.02, s = 0.45, h/r = 1, s = 0.05

h/r=5

3.5

=0
0.10
0.20
0.30
0.60
1

0.1

2.5

s = 0
0.05
0.20

si

2.0

s = 0
0.05
0.20

1.5

1.0
0.2

0.4

0.6

0.8

0.01

1/

1.0

0.1

1/

(d) = 0.02, s = 0.45, h/r = 5, = 0.15


(c) = 0.02, s = 0.45, = 0.15
Fig: 4. Parametric results using the proposed exact procedure. (a) Period of SSI system as function of , (b)
Effective damping of SSI system as function of , (c) Period of SSI system as function of soil damping
ratio s, (d) Effective damping of SSI system as function of s

194

2.5

1.0
0.9
0.8
0.7

2.0

0.6
0.5
0.4

1.5

=0
0.20
0.50
0.80
1

1.0
0.2

(a)

0.4

0.6

0.8

1.0

=0
0.20
0.50
0.80
1

0.3
0.2
0.1

1/

0.0
0.2

= 0.02, s = 0.45, h/r = 0.33, s = 0.05

(b)

2.5

0.4

0.6

0.8

1.0

1/

= 0.02, s = 0.45, h/r = 0.33, s = 0.05

0.14

0.12

2.0

0.10

0.08

=0
0.20
0.50
0.80
1

1.5

0.06

=0
0.20
0.50
0.80
1

0.04

0.02

1.0

(c)

1/
0.4
0.6
0.8
1.0
= 0.02, s = 0.45, h/r = 2, = 0.15, s =0.05

0.00

0.2

(d)

1/
0.2
0.4
0.6
0.8
1.0
= 0.02, s = 0.45, h/r = 2, = 0.15, s= 0.05

Fig 5:. Influence of foundation mass on the dynamic behavior of SDOF systems

i xx
K

i xr
K
i rr
K

I=

ix
K

i
K

Ep I p
d

(a)

(c)

(b)

Fig 6:. (a) Model of pile-supported-structure, (b) Distribution of soil reactions due to horizontal loading, (c)
Reduced model with two dynamic impedances

195

3.0

0.30

0.25

h/d = 1
5
10

2.5

0.20

2.0

0.15

0.10

1.5

h/d = 1
5
10

0.05

1.0
0.2

0.4

0.6

0.8

1.0

1/

0.00
0.2

(a) = 0.02, s = 0.45, = 0.15, s = 0.05, p/s = 1.40, Ep/Es = 100

0.6

0.8

1.0

1/

(b) = 0.02, s = 0.45, = 0.15, s = 0.05, p/s = 1.40, Ep/Es = 100

2.5

0.12

Ep/Es = 100
1000
10000

0.10

2.0

0.4

0.08

i
1.5

0.06

0.04

Ep/Es = 100
1000
10000

0.02

1.0
0.2

0.4

0.6

0.8

1.0

1/

0.00
0.2

(c) = 0.02, s = 0.45, = 0.15, s = 0.05, p/s = 1.40, h/d = 5

0.4

0.6

0.8

1.0

1/

(d) = 0.02, s = 0.45, = 0.15, s = 0.05, p/s = 1.40, h/d = 5

Fig 7:. (a) System period as function of h/d, (b) System damping as function of h/d. (c) System period as
function of Ep/Es, (d) System damping as function of Ep/Es
0.12

0.28
0.24

0.10

Pile
Footing

0.20

Pile
Footing

0.08

0.16

0.06
0.12

0.04

0.08
0.02

0.04
0.00
0.1

0.2

0.3

0.4

(f)

(a) = 0.02, s = 0.45, h/d = 1, s = 0.05, p/s = 1.40, Ep/Es = 1000

0.00
0.02

0.08

0.12

0.16

(f)

(b) = 0.02, s = 0.45, h/d = 5, s = 0.05, p/s = 1.40, Ep/Es = 1000

(p)

1.0

0.04

1
(1/)
f

0.20

0.8
0.15

0.6

Pile
Footing

Pile
Footing
0.10

0.4
(f)

0.2

0.05

0.0
0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

0.00

(a0)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

(a0)

(c) s = 0.45, s = 0.05, p/s = 1.40, Ep/Es = 1000


(d) s = 0.45, s = 0.05, p/s = 1.40, Ep/Es = 1000
Fig 8: (a, b) Effective damping of a structure founded on pile and footing, (c) Radiation damping of a pile and
footing due to translational oscillation, (d) Radiation damping of a pile and footing due to rocking
oscillation

196

Interaction of Shallow and Deep Foundations


with a Rupturing Normal Fault
I. Anastasopoulos and G. Gazetas
National Technical University of Athens, Greece

Abstract
The paper studies the response and distress of shallow and deep foundations to a seismic
fault rupture emerging directly underneath. The developed numerical methodology has been
calibrated with centrifuge experiments. The outlined parametric results provide valuable
insight to the respective soilfoundation interplay, and could explain qualitatively the
observed behaviour in a number of case histories from recent earthquakes. It is shown that
rigid mat or box foundations may divert the rupture path and, if properly designed, survive
the rupture with only some unavoidable rotation. The role of piles is not nearly as clear, and
the paper highlights a possibly detrimental effect.

INTRODUCTION

field studies, centrifugal experiments, and


numerical / analytical modeling. Specifically:
Field studies of documented case
histories motivated our investigation and
offered material for calibration of the
theoretical methods and analyses,
Carefully
controlled
centrifugal
experiments helped in developing an
improved understanding of mechanisms
and in acquiring a reliable experimental
data base for validating the theoretical
simulations, and
Theoretical
methods
(analytical
or
numerical) calibrated against the above
field and experimental data offered
additional insight into the nature of the
interaction,
and
were
utilised
in
developing parametric results and design
aids.
This paper summarises some of the key
findings of these theoretical studies, which were
later supplemented with further analyses
pertaining to pile and caisson foundations.

Numerous cases of devastating effects of


earthquake surface fault rupture on structures
were observed in the 1999 earthquakes of
Kocaeli, Duzce, and Chi-chi. However,
examples of satisfactory, even spectacular,
performance of a variety of structures also
emerged. In some cases the foundation and
structure were quite strong and thus either
forced the rupture to deviate or withstood the
tectonic movements with some rigid-body
rotation and translation but without damage. In
other cases structures were quite ductile and
deformed without failing. Thus, the not-so-new
idea that a structure can be designed to survive
with minimal damage a surface fault rupture reemerged. The work presented here was
motivated by the need to develop deep
quantitative understanding of the interaction
between a rupturing normal fault and a variety
of foundation types. Much of the reported
research was conducted within the framework
of
a
Europian-Union
research
project
(QUAKER) with the participation of the
research teams from the University of Dundee,
Godynamique
et
Structure,
Studio
Geotechnico Italiano, and National Technical
University of Athens.
The study was based on an integrated
approach, comprising three interrelated steps:

FAULTRUPTURE PROPAGATION AND


INTERACTION WITH FOUNDATIONS

ITS

Statement of the Problem


It has long been recognised (Duncan &
Lefebvre, 1973; Bray 1990) that a strong

197

structure founded on/in soil can resist


successfully the loading induced by a rupturing
seismic fault. In the Kocaeli, Dzce-Bolu, and
Chi-Chi earthquakes of 1999 numerous
structures (single-storey and multi-storey
buildings, bankers, bridge piers, retaining
structures, electricity pylons, dams, tunnels)
were located directly above the propagation
path of the rupturing (normal, strike-slip,
reverse) faults. Some of these structures
exhibited a remarkably good behaviour. This
observation had a strong motivating influence
for our research effort.
For it became
immediately clear that the strict prohibition: Do
not build in the immediate vicinity of active
faults, which the prevailing seismic codes
invariably imposed, was unduly restrictive (and
in many cases meaningless).
Indeed, along the ground surface in the
freefield, ruptures are neither continuous, nor
do they follow precisely the surface outcrop of
pre-existing faults (Ambraseys & Jackson,
1984).
In addition to several geologic factors that
contribute to such behaviour, significant
appears to be the role of a soil deposit that
happens to overlie the rock base through which
the rupture propagates. If, where, and how
large will the dislocation emerge on the ground
surface (i.e. the fault will outcrop) depends not
only on the style and magnitude of the fault
rupture, but also on the geometric and material
characteristics of the overlying soils. Field
observations and analytical and experimental
research findings [Bray et al, 1994a, b; Cole &
Lade, 1984; Lade et al, 1984; Lazarte & Bray,
1995] show that deep and loose soil deposits
may even mask a small-size fault rupture which
occurs at their base; whereas by contrast with a
cohesive deposit of small thickness, a large
offset in the base rock will likely cause a distinct
fault scarp of nearly the same displacement
magnitude. One important finding of the above
studies is that the rupture path in the soil is not
a simple extension of the plane of the fault in
the base rock: phenomena such as diffraction
and bifurcation affect the direction of the
rupture path, and make its outcropping location
and offset magnitude difficult to predict.
Our interest here is not on the propagation
of a rupture within the soil, but on how a
structure sitting on top of the fault breakout
behaves. It turns out that a fascinating interplay

takes place between the propagating fault


rupture, the deforming soil, the differentially
displacing foundation, and the supported
structure. Two different phenomena take place.
First, the presence of the structure modifies the
rupture path. Depending on the rigidity of the
foundation and the weight of the structure, even
complete diversion of the fault path before it
outcrops may take place. Obviously, the
damage to a given structure depends not only
on its location with respect to the fault outcrop
in the free-field, but also on whether and by
how much such a diversion may occur.
Second, the loads transmitted from the
foundation on to the soil tend to compress the
asperities and smoothen the anomalies of
the ground surface that are produced around
the fault breakout in the freefield, i.e. when the
structure is not present. Thus, depending on
the relative rigidity (bending and axial) of the
foundation with respect to the soil, as well as on
the magnitude of the structural load, the
foundation and the structure will experience
differential displacements and rotation different
from those of the free-field ground surface.
This phenomenon, given the name FaultRuptureSoilFoundationStructure Interaction
[FRSFSI] by Anastasopoulos & Gazetas
(2007), is briefly elucidated in the sequel for
shallow and deep foundations.
Numerical Analysis and Results: Shallow
Foundations
The problem studied here is illustrated in Fig. 1.
We consider a uniform soil deposit of thickness
H at the base of which a normal fault, dipping at
an angle (measured from the horizontal),
produces
downward
displacement
(dislocation, offset) of vertical amplitude h.
The analysis is conducted in two steps. First,
fault rupture propagation through soil is
analysed in the free field, ignoring the presence
of the structure. Then, a strip foundation of
width B carrying a uniformly distributed load q
or a multistory framestructure is placed on top
of the freefield fault outcrop at a specified
distance s (measured from its corner), and the
analysis of deformation of the soilstructure
system due to the same base dislocation h is
performed. The analyses are conducted under
2-D plane-strain conditions evidently a
simplification, in view of the finite dimensions of

198

B
q

x
y
s

Hanging wall

free-field
rupture path

H H

Footwall

Figure 1: Configuration of the soilfoundation system subjected to a normal fault dislocation at the base rock.

detachment of the foundation from the bearing


soil (i.e. gap formation beneath the foundation).
The interface shear properties follow Coulombs
friction law, allowing for slippage. Both
detachment and slippage are important
phenomena for a realistic foundation model.
A typical result elucidating the interplay
between loose (Dr = 45%) soil, rupture path,
and a perfectly rigid foundation carrying a 4storey structure is given in Fig. 2. A base rock
dislocation of 2 m (5% of the soil thickness) is
imposed.
The
structure
is
placed
symmetricallystraddling the free-field fault
breakout (i.e. the foundation is placed with its
middle coinciding with the location where the
fault would outcrop in the free field). Yet, a
distinct rupture path (with high concentration of
plastic shearing deformation and a resulting
conspicuous surface scarp) is observed only in
the freefield. The presence of the structure
with its rigid foundation causes the rupture path
to bifurcate at about the middle of the soil layer.
The resulting two branches outcrop outside the
left and the right corner of the foundation,
respectively. The soil deformations around
these branches are far smaller and diffuse than
in the freefield, and the respective surface
scarps are much milder. Thanks to the
substantial weight of the structure and the
flexibility of the ground, the structure settles and
rotates as a rigid body. The foundation does
not experience any loss of contact with the
ground ; apparently, the foundation pressure is
large enough to eliminate any likely asperities
of the ground surface.
As a result of such behaviour, the structure
and its foundation do not experience any

a real structure in the direction parallel to the


fault. The relative location of outcropping is
varied parametrically through the distance S.
Comparing
soil
and
groundsurface
deformations in the two steps gives a first
picture of the significance of SFSI.
Among several alternatives that were
explored, the FE model shown in Fig. 2
produced results in excellent accord with
several centrifugal experiments conducted at
the University of Dundee for both steps of the
analysis (Anastasopoulos et al 2007 a, b). A
parametric investigation revealed the need for a
long (B = 4H) and very refined mesh (element
size of 0.5 m 1.0 m) along with a suitable slipline tracing algorithm in the region of soil
rupture and foundation loading. An elastoplastic
constitutive model with the Mohr-Coulomb
failure criterion and isotropic strain softening
was adopted and encoded in the ABAQUS
finite element environment. Similar models
have been successfully employed in modeling
the failure of embankments and cut slopes
(Potts et al, 1987, 1990). Modelling strain
softening was shown to be necessary; it was
introduced by suitably reducing the mobilised
friction angle mob and the mobilised dilation
angle mob with increasing plastic octahedral
shear strain. With all the above features, the FE
formulation is capable of predicting realistically
the effect of large deformations with the
creation and propagation of shear bands.
The foundation, modeled with linear elastic
beam elements, is positioned on top of the soil
model and connected to it through special
contact elements. The latter are rigid in
compression
but
tensionless,
allowing

199

L = 4H
free-field
fault outcrop

Foot wall

(a)

(b)

Hanging wall

Foundation

Figure 2:

Figure 2. Finite element discretisation and the two steps of the analysis: (a) fault rupture
propagation in the free-field, and (b) interplay between the outcropping fault rupture and the
structure (termed Fault RuptureSoilFoundationStructure Interaction, FR-SFSI).

substantial distress, while their rotation and


settlement could perhaps be acceptable.

and kinematic () characteristics of the soil


along the depth.

The main factors influencing FR-SFSI are:

The style of faulting (normal, thrust, strikeslip), the angle of dip and the
offset (dislocation) at the basement rock.
The total thickness (H) of the overlying soil
deposit, and the stiffness (G), strength (, c)

200

The type of the foundation system (for


example, isolated footings, mat foundation,
box-type foundation, piles, caissons).

The flexural and axial rigidity of the


foundation system (thickness of mat
foundation cross-section and length of tie
beams, etc.)

The load of the superstructure and the


foundation.

The stiffness of the superstructure (cross


section of structural members, spacing of
columns, presence or not of shear walls).

he location s from the foundation corner to


the free-field outcrop.

dislocation h/H = 5% would cause only a minor


diversion of the rupture path, easily noticeable
only in the loose soil (about 2 m towards the
hanging wall, to the left of the foundation). The
differential settlement is higher on loose sand.
The main difference between the two soils is in
the uplifting of the foundation. In dense sand
the building loses contact at both sides, uL 5
m, uR 3 m ; with only its central part
maintaining contact over a width bC 12 m. On
the other hand, in loose sand the building uplifts
only at the left side, uL 3 m.
The foundation distress is about 50% higher
in the dense sand, as a result of the creation of
a wider cantilever, whereas in the loose sand
the greater compression of the scarp is
beneficial. However, the differential settlement
and (rigid-body) rotation of the foundation is
three-time higher on the loose sand.

However, a detailed investigation of the role


of all the above parameters is beyond the
scope of this chapter. Reference is made to
Anastasopoulos (2005) and Anastasopoulos &
Gazetas (2007, b) for such a parameter study.
Here we only outline a few characteristic results
pertaining to a 20 m wide rigid mat foundation,
transmitting 30 kPa uniform pressure. The soil
layer is either loose (Dr 45%) or dense (Dr
80%) sand of total thickness H = 40 m. Three
locations of the foundation with respect to the
free-field outcrop are considered: s = 4 m,
10 m, and 16 m, i.e. near the left edge, in the
middle, and near the right edge of the
foundation, respectively.
Figs 35 portray the response of the
soilfoundation system for each location S and
each of the two soil densities, for a
parametrically variable ratio of base dislocation
over layer thickness:

.
Shown in each figure are the deformed
mesh, the distribution of plastic strains, the
diversion of the rupture D, the vertical
displacement profile y, the distortion angle ,
and the contact pressures p along the soilfoundation interface. In all cases the results are
compared with the corresponding free-field
results, to visualize the effects of FR-SFSI. The
contact stresses are compared to their initial
distribution (i.e., for h/H = 0, before the bedrock
displacement is applied) to reveal which parts of
the structure are losing contact with the bearing
soil, and hence foundation uplifting takes place.
The left part of the building that uplifts will be
denoted as uL, the right uR, and uC if the uplifting
takes place around the centre. In similar
fashion, the part of the foundation that
maintains contact will be denoted as bL, bR and
bC, if it is located at the left side, the right side,
or the middle, respectively.
The following trends are worthy of note:

(2) For the fault emerging (without SFSI) in


the middle of the foundation (s = 10 m, or s/B =
0.50) the rupture path is diverted, becomes very
diffuse, and bifurcates for h/H = 5%. The left
branch (which is a secondary one) diverts by
about 3 m to 4 m towards the hanging wall. In
both cases of loose and dense sand a fault
scarp develops beneath the building. The
foundation maintains always contact at its left
edge and at its middle part. In dense sand,
moving from left to right, there is first a small
part of the building bL 2 m that is in contact,
followed by an uplifted portion, uL 4 m, then
the middle part that remains in contact, bR 10
m, and finally the far most right part of the
foundation that uplifts uR 4 m. Although the
situation is qualitatively similar with loose sand,
uplifting is much less extensive (uR 1 m). The
effective width of the foundation, i.e. in contact
with the soil, is now: bL = bC 19 m. And the
differential settlement is about 2 m in both
cases.
(3) For the fault emerging (without SFSI)
near the right edge of the foundation (s = 16 m,
or s/B = 0.80), we see again diversion,
diffusion, and a minor bifurcation of the rupture
path for h/H = 5%. The right branch (which is
the most significant) diverts slightly towards the
footwall by about 24 m, while the (barely
noticeable) left branch is diverted towards the
hanging wall ( 3 m). But a clear significant
difference is noted between the loose and
dense sand cases:

(1) For the fault emerging (without SFSI)


near the left edge (s = 4m, or s/B = 0.20), this
lightly loaded foundation for a relative base

201

Dense Sand

Loose Sand
(i)

Dense
Sand
B = 20
m

(i)

Loose
Sand
B = 20
m

y = 21 c m

y = 61 c m

q = 30 kPa

q = 30 kPa

Free Field
Free Field

foundation
building

foundation
building

0
h/H = 1 %

h/H = 1 %

h/H = 3 %

-1

(ii)

-2
10

15

20

25

h/H = 5 %

(ii)
0

30

10

15

20

25

30

foundation
building

-50

-50

pv (kPa)

pv (kPa)

h/H = 4 %

-2

foundation
building

h/H = 0 %

-100

-150

h/H = 0 %

-100

h/H = 5 %

-150
h/H = 5 %

(iii)

(iii)

-200

-200
0

10

80

15

20

25

30

10

80

foundation
building

60

60

40

40

(%)

(%)

-1

-1.5

h/H = 5 %

h/H = 3 %

free field

h/H = 4 %

-1.5

h/H = 2 %

-0.5

h/H = 2 %

y (m)

y (m)

-0.5

20

15

20

25

30

foundation
building

free field

20
0

(iv)

-20
0

10

15

20

25

(iv)

-20
0

30

Di stance (m)

10

15

20

25

30

Distance (m)

Figure 3: FR-SFSI analysis of rigid B = 20 m foundation subjected to q = 30 kPa surcharge load. Fault
rupture in the free field emerging at s = 4 m : (i) Deformed mesh and plastic strain, (ii) Vertical
displacement at the surface, (iii) contact pressure p, and (iv) distortion angle . The results of the
FR-SFSI analysis (red lines) are compared with the Free-field results (blue lines) for h/H = 1
to 5%.

202

Dense Sand

Loose Sand
(i)

B = 20
m
Dense
Sand

(i)

Loose
B = Sand
20 m

y = 218 c m

q = 30 kPa

y = 191 c m

q = 30 kPa

Free Field
Free Field

foundation
building

foundation
building

h/H = 1 %

h/H = 1 %

-0.5

h/H = 2 %

y (m)

y (m)

-0.5

h/H = 3 %

-1

h/H = 4 %

-1.5

h/H = 2 %
h/H = 3 %

-1

h/H = 4 %

-1.5

h/H = 5 %

h/H = 5 %

free field

-2

-2

(ii)
0

10

15

20

25

30

building
foundation

10

15

20

25

30

foundation
building

-50

pv (kPa)

pv (kPa)

-50
h/H = 0 %

-100

-150

-100

h/H = 0 %

h/H = 5 %

-150
h/H = 5 %

(iii)

-200
0

10

15

20

25

(iii)

-200

30

80

10

80

foundation
building
60

15

20

25

30

foundation
building

60

40

40

free field

(%)

(%)

(ii)

-2.5

-2.5

20

20

-20

-20

(iv)

-40
0

10

15

20

25

(iv)

-40
0

30

Di stance (m)

10

15

20

25

30

Di stance (m)

Figure 4: FR-SFSI analysis of rigid B = 20 m foundation subjected to q = 30 kPa surcharge load. Fault
rupture in the free field emerging at s = 10 m : (i) Deformed mesh and plastic strain, (ii) Vertical
displacement at the surface, (iii) contact pressure p, and (iv) distortion angle . The results of the
FR-SFSI analysis (red lines) are compared with the Free-field results (blue lines) for h/H = 1
to 5%.

203

Dense Sand

Loose Sand

B = 20 m

(i)

(i)

B = 20 m

y = 150 cm

y = 82 cm

q = 30 kPa

q = 30 kPa

Free Field
Free Field

foundation
building

foundation
building

0
h/H = 1 %

h/H = 1 %

-0.5

h/H = 2 %

-1

y (m)

y (m)

-0.5

h/H = 3 %
h/H = 4 %

-1.5

h/H = 4 %

-2

(ii)

-2.5

(ii)
-2.5

10

15

20

25

30

foundation
building

10

15

20

25

30

foundation
building

h/H = 0 %

h/H = 0 %

-100

pv (kPa)

-100

-200

h/H = 5 %

-200

-300

-300
h/H = 5 %

(iii)

-400
0

10

15

20

25

(iii)

-400
0

30

10

15

20

25

30

120

120

foundation
building

foundation
building
80

(%)

80

(%)

h/H = 3 %

h/H = 5 %

-2

pv (kPa)

-1

-1.5

free field

h/H = 5 %

h/H = 2 %

free field

40

40

(iv)

-40
0

10

15

20

25

(iv)

-40
0

30

Di stance (m)

10

15

20

25

30

Di stance (m)

Figure 5: FR-SFSI analysis of rigid B = 20 m foundation subjected to q = 30 kPa surcharge load. Fault
rupture in the free field emerging at s = 16 m : (i) Deformed mesh and plastic strain, (ii) Vertical
displacement at the surface, (iii) contact pressure p, and (iv) distortion angle . The results of the
FR-SFSI analysis (red lines) are compared with the Free-field results (blue lines) for h/H = 1
to 5%.

204

The fault scarp that is formed near the


right edge of the foundation is
conspicuous only with loose sand.
On dense sand, the middle part of the
foundation loses contact with the bearing
soil, uC 11 m, while the left and right part
of it remain in contact, bL 2 m and bR
7 m.
On loose sand, the response is quite
favourable: not only is the dislocation
diverted by more than 4 m and outcrops
beyond the right edge of the structure, but
full contact is maintained over the whole
length of the soilfoundation interface.
The distress of the foundation is thus
significantly less with loose than with
dense sand. Also smaller on loose sand
is the (rigid-body) rotation of the
foundation.
Such a good response of a building on loose
soil on the hanging wall is reminiscent of
several success stories from the Kocaeli 1999
earthquake, especially of the building in
Denizerler across the entrance from the Ford
factory, near Glck (see Anastasopoulos et
Gazetas 2007).
(4) Although not shown here, the effect of an
increase in the transmitted load from 30 kPa to
60 kPa is quite beneficial on loose sand, but
almost negligible on dense sand. The most
significant benefits are the decrease of
foundation rotation (and of structure tilting) and
the elimination of a large part of uplifting. As a
consequence, the survival of a heavy
structure on top of a major fault rupture in loose
soil seems quite possible, in qualitative accord
with numerous such success stories in several
earthquakes.

tied to the different blocks of the fault may


indeed be vulnerable. An interesting analogy
has been brought to our attention by Professor
J. Bray (2005): deep-rooted trees being torn
apart by a fault rupturing directly underneath,
apparently as a result of their roots being pulled
in opposite directions.
Two typical foundation systems are examined
here in order to highlight the interaction
between a deep foundation and an emerging
fault rupture:
a 3x3 capped pile group
a square rigid embedded foundation
(caisson)
A 3-D finite element model was developed for
each case, using eight-noded elements, and
employing the same soil and interface
constitutive models as described in the
preceding section. Fig 6 presents a plane
section of the complete model. In both cases
the soil deposit consisted of dense sand, of
total thickness H = 20 m. Needless to say, the
choise of this limited depth was motivated solely
by the desire for the smallest possible size of
this 3D model.

Numerical Analysis
Foundations

and

Results:

Piles
The piles are of length Lp = 15 m, diameter
dp = 1 m and are spaced 4 m apart (from axis
to axis). Their cap is 10 m x 10 m in plan and
2.5 m thick, and carries a structural vertical load
of 10 MN. A rigid connection is assumed
between cap and piles (fixed-head piles). Only
ideally elastic pile behaviour is considered at
the present time, although the necessity for
accounting for pile inelasticity will become
apparent (if a realistic assessment of the
response of the system to large fault offsets is
needed).
Aiming at giving a first picture of the
possible straining to be experienced by the
piles, Fig. 7 portrays the deformed finiteelement mesh with the distribution of plastic
shear strains. Four positions of the pile group
with respect to the outcropping fault in the
freefield are examined: S = 1, 5, 9, and 13
meters, where S is measured from the edge of
the pile cap (which lies 1 m to the left of the
nearby pile axis). Then Fig. 8 presents detailed
results (deformations and internal forces) for the
case of S = 5 m, only. Several trends are worth
noting in these figures:

Deep

Whereas piles are used for protecting structures


by helping to keep total and differential
settlements small, their role in supporting
structures straddling seismic faults is far from
clear. Scant (perhaps only circumstantial)
evidence from recent earthquakes has
implicated the piles in some structural damage
see for example the analysis of the damage
of the pile-supported Attaturk Stadium in
Denizerler during the Kocaeli Earthquake
(Anastasopoulos & Gazetas, 2007). Systems

205

a
Axis of symmetry

Free field rupture path

h
b

a
Axis of symmetry

Free field rupture path

Figure 6: Pile group and caisson foundations in the path of a rupturing fault. Cross section aa of the 3D finite
discretisation

larger than the imposed base offset, .


The pile group, however, remains almost
intact: there is no displacement or rotation
of the pile and only the piles in the front
row experience some (rather minor)
distress (bending moments of the order of
300 kNm).
(2) For s = 5 m, the fault would have emerged
at the center of the foundation in the

(1) For s = 1 m, when the fault emerges near


the left edge of the pile group (and the
group is therefore almost all in the
footwall), a slight diversion of the rupture
path to the left takes place. A very distinct
scarp is formed immediately next to the
piles. The scarp forms a slope of about the
same angle, a, as that of the triggering
basement rupture, and is appreciably

206

(3) For s = 9 m, the fault would have emerged


near the right edge of the group in the free
field. The piles with their presence and
transmitted loads diffuse the rupture,
thereby suffering unequal settlements and
non-uniform large displacements. As a
result,
the
rotation
and
lateral
displacement of the cap and the bending
moments in the piles attain very large
(unacceptable) values.

freefield. The presence of the axially


loaded piles makes the rupture path: (i) to
partly divert to the left and emerge just at
the edge of the front of piles, and (ii) to
become diffuse in the region between the
piles. Substantial rotation and horizontal
displacement of the pile cap take place.
The front row of piles is being pulled
outward and downward by the dropping
hanging wall of the fault; as a result very
large bending moments would develop at
the pilehead, in excess of 12 MNm for a
dislocation of 2 m. The middle row of piles
would experience much less distress, but
the last row and especially the corner piles
would
develop
substantial
bending
moments (almost 6 MNm for a dislocation
of 2 m), as a result of being pushed near
their middle. Notice the completely
different pattern of bending moments with
depth between front-row and back-row
piles in Fig. 7 : whereas for pile 1 (front
row) the maximum is at the top, for pile 6
(corner pile) the maximum appears at 10
m depth. Evidently, such large bending
moments, especially in the front row,
exceed the maximum conceivable capacity
of a well-reinforced 1 m diameter pile,
implying structural failure at least of
conventional-type piles.

(4) Finally, for s = 13 m (freefield fault outcrop


3 m beyond the pile-cap), there is a slight
diversion of the rupture to the right with a
simultaneous slight diffusion of the plastic
shear strains. Only the last row of piles is
stressed significantly (max M 4MNm for
h = 2 m). There is apparently no rotation of
the pile cap, but a downward and outward
displacement are unavoidable; such
displacements might have a detrimental
effect on a framed structure, one column of
which is supported on the studied piledfoundation.
In conclusion, it appears that the response of
piled foundations may be less favourable than
that of rigid mat foundations. However, two
significant limitations of the performed analyses
on which these conclusions are partly based
must be noted here:

s=1 m

s=5 m

s=9 m

s = 13 m

Figure 7: Deformed mesh of the soilpilecap system with the concentration of plastic octahedral strains, for
different positions (s = 1 13 m) of the emerging fault rupture.

207

vertical component of dislocation h = 2 m.


By contrast, recall that the corresponding
piled foundation had developed a rotation
and horizontal displacement, while its front
row of piles had been substantially
distressed.
Notice also that a secondary rupture has
begun to form, propagating at an angle of
about 30o to the left of main rupture. It is
about to reach the ground surface for h = 2
m, and the associated graben between the
two normal ruptures is (barely) visible in the
scale of the figure.

perfect (bonded) contact was assumed


between piles and soil
the piles were modelled as a perfectly
elastic material.
As a result of the first assumption, the forces
upon the piles by the outward and downward
moving hanging wall are exaggerated. Soil
sliding around the piles would reduce the
magnitude of such drag forces, thereby
leading to smaller pile distress and smaller cap
rotation / displacement. Regarding the second
assumption, note that the large bending
moments in the piles would not of course
materialize in reality, since their ultimate
structural capacity can not be exceeded.
Prediction of the consequences of the
unavoidable redistribution of loads among the
piles, and among piles and raft, can not be
made reliably with the results presented above
for purely elastic piles.

(3) For s = 9 m, the rupture is diffused the


caisson rotates substantially to the left (8o
for h = 2 m), and an active state of stress
develops on the back side of the caisson.
Clearly this behaviour is not so favourable;
for instance it would cause distress in a
framed structure one column of which is
supported on such a caisson. But by
contrast to the piled foundation, the
capability of the caisson to transmit the
vertical load would be hardly affected.

Rigid Caisson
The caisson is 10 m x 10 m in plan and also 15
m in depth. It carries 10 MN vertical load. Only
fully bonded contact between the caisson and
the soil is considered --- an idealization that is
likely to lead to a conservative assessment of
the caisson displacement / rotation.
Dominant role in the response of a given
caisson to fault rupturing underneath plays its
position with respect to the freefield rupture
outcropping. Again four such positions are
considered: 1, 5, 9, and 13 meters. Figs 8-9
portray the deformed mesh with the distribution
of plastic octahedral shear strains for each
value of s. Fig. 8 gives the plane section
(along the axis) while Fig. 9 depicts a 3-D view
(of half the model). The following conclusions
are drawn:

(4) For s = 13 m, the rupture path hits the


base corner of the caisson and defracts to
the right, emerging at the ground surface at
a distance of S = 18 m, i.e., 5 m to the right
of the freefield outcrop. The caisson
essentially follows the movement of the
hanging wall, thereby experiencing an
appreciable rotation of about 3o for h = 2 m.
In conclusion, it appears that the response of
deep embedded foundations (caissons) would
in most cases be quite satisfactory, especially if
structural provisions are taken to accommodate
their unavoidable rotation at large fault offsets.
Once again, one of the limitations in our
modelling, namely the assumption of a
perfectly-bonded
interface,
may
have
exaggerated the lateral displacement/rotation of
the caisson.

(1) For s = 1 m, the fault emerges to the left of


the caisson, diverted slightly, and forms a
distinct scarp similar to that in the case of
the piled foundation. The caisson does not
experience any measurable rotation or
displacement.

ACKNOWLEDGEMENTS
This work formed part of the EU research
project QUAKER, funded through the EU Fifth
Framework Programme: Environment, Energy,
and Sustainable Development, Research and
Technological Development Activity of Generic
Nature: the Fight against Natural and
Technological Hazards (contract number:

(2) More significant is the diversion of the


rupture path in case of s = 5 m; the fault
now emerges vertically along the side of the
caisson. The latter hardly feels the rupture,
experiencing a rotation of merely 1o for a

208

aa

bb

pile cap

pile cap

0
h=0.2 m

h=0.2 m
h=0.4 m
h=0.8 m

-1
h=1.2 m

-1.5

h=0.8 m

-1
h=1.2 m

-1.5

h=1.6 m

-2

h=0.4 m

-0.5

y (m)

y (m)

-0.5

h=1.6 m

-2

h=2 m

x (m)

-2.5
-10

10

h=2 m

x (m)

-2.5
-10

20

10

20

16

16

Pile 2

Pile 1

12

Pile 5

M (MNm)

M (MNm)

12

-5

0.5

1.5

h (m)

M (MNm)

M (MNm)

10

15

-5
0

static

y (m)

y (m)

0.5

h (m)

1.5

10

15

Pile 6
static
h=0.1

h=0.1

12

h=0.8

16

Pile 4

0
0

Pile 1

12

Pile 6

Pile 3

h=2.0

16

h=0.8
h=2.0

Figure 8: Detailed results (vertical displacement y of the ground surface, largest bending moment in the
piles, and distribution of bending moments in piles 1 and 6 for a 3 x 3 capped pile group, for a fault
rupturing position s = 5 m.

209

s=1 m

s=5 m

s=9 m

s = 13 m

s=1 m

s=5 m

s=9 m

s = 13 m

(a)

(b)

Figure 9: Deformed mesh of the caissonsoil system with the concentration of plastic octahedral strains, for
different positions (s = 1 13 m) of the emerging fault rupture: (a) Section aa ; (b) 3-D view.

REFERENCES AND BIBLIOGRAPHY

EVG1-CT-2002-00064). The research on pile


and caisson foundations built on top of a
rupturing normal fault was part of a project
funded by the Greek Railway Organization
OSE.

Anastasopoulos, . (2005), Fault RuptureSoil


FoundationStructure
Interaction,
Ph.D.
Dissertation, School of Civil Engineering,
National Technical University, Athens, pp.570.

210

Anastasopoulos I., Gazetas G., Bransby M.F.,


Davies M.C.R., and El Nahas A.(2007), "Normal
Fault Rupture Interaction with Strip Foundations",
Journal of Geotechnical and Geoenvironmental
Engineering, Vol. 133

Journal of the Soil Mechanics and Foundation


Division, ASCE, Vol. 99, pp. 1153-1163.
Lade, P.V., Cole, D.A., Jr., and Cummings, D.
(1984), Multiple Failure Surfaces Over Dip-Slip
Faults, Journal of Geotechnical Engineering,
ASCE, Vol. 110, No. 5, pp. 616-627.

Anastasopoulos I., & Gazetas G. (2007),


"Foundation-Structure Systems over a Rupturing
Normal Fault : Part I. Observations after the
Kocaeli
1999
Earthquake",
Bulletin
of
Earthquake Engineering, Vol. 5, No. 3

Lazarte, C.A., and Bray, J.D. (1995), Observed


Surface Breakage due to Strike-Slip Faulting,
Third International Conference on Recent
Advances in Geotechnical Engineering and Soil
Dynamics, Vol. 2, pp. 635640.

Anastasopoulos I., & Gazetas G. (2007), "Behaviour


of Structure-Foundation Systems over a
Rupturing Normal Fault : Part II. Analysis of the
Kocaeli Case Histories", Bulletin of Earthquake
Engineering, Vol. 5, No. 3

Potts, D. M., Dounias, G. T. & Vaughan, P. R. V.


(1987), Finite element analysis of the direct
shear box test, Gotechnique, Vol. 37, No. 1, pp.
1123.

Anastasopoulos I., Gazetas G., Bransby M.F.,


Davies M.C.R., and El Nahas A.(2007), "Fault
Rupture Propagation through Sand : Finite
Element Analysis and Validation through
Centrifuge
Experiments",
Journal
of
Geotechnical
and
Geoenvironmental
Engineering , Vol. 133

Potts, D. M., Dounias, G. T. & Vaughan, P. R. (1990),


Finite element analysis of progressive failure of
Carsington Embankment, Gotechnique, Vol. 40,
No. 1, pp. 79101.
Potts, D. M., Kovacevic, N. & Vaughan, P. R. (1997),
Delayed collapse of cut slopes in stiff clay,
Gotechnique, Vol. 47, No. 5, pp. 953982.

Ambraseys, N. & Jackson, J., (1984), Seismic


Movements, Ground Movements and their
Effects on Structures, P.B. Attewell and R.K.
Taylor Editors, Surrey University Press, pp. 353380.
Bray, J.D. (1990), The effects of tectonic movements
on stresses and deformations in earth
embankments, Ph.D. Dissertation, University of
California, Berkeley.
Bray, J.D. (2001), Developing mitigation measures
for the hazards associated with earthquake
surface fault rupture, Workshop on Seismic
Fault-Induced Failures: Possible Remedies for
Damage to Urban Facilities, Japan Society for
the Promotion of Science, University of Tokyo,
pp. 5579.
Bray, J.D., Seed, R.B., Cluff, L.S., and Seed, H.B.
(1994), Earthquake Fault Rupture Propagation
through
Soil,
Journal
of
Geotechnical
Engineering, ASCE, Vol. 120, No.3, March, pp.
543-561.
Bray, J.D., Seed, R.B., and Seed, H.B. (1994b),
Analysis
of
Earthquake
Fault
Rupture
Propagation through Cohesive Soil, Journal of
Geotechnical Engineering, ASCE, Vol. 120, No.3,
March, pp. 562-580.
Cole, D.A. Jr., and Lade, P.V. (1984), Influence
Zones in Alluvium Over Dip-Slip Faults, Journal
of Geotechnical Engineering, ASCE, Vol. 110,
No. 5, pp. 599-615.
Duncan, J.M., and Lefebvre, G. (1973), Earth
pressures on Structures Due to Fault Movement,

211

The Role of the Soil Stiffness on the Dynamic Impedance Functions


D. Pitilakis1, D. Clouteau2, A. Modaressi2
1

Ecole Centrale Paris, France (currently in Aristotle University of Thessaloniki, Greece)


2
Ecole Centrale Paris, France

Abstract
This paper provides an insight in the role of the soil stiffness on the estimation of the
dynamic impedance functions. The softening of the soil under strong ground shaking is not
taken into account in the dynamic impedance functions, under the assumption of linear
elastic soil behavior. Nevertheless, the resulting shear wave velocity reduction due to
nonlinear soil behavior may have important effects on the amplitude and shape of the
dynamic stiffness and radiation damping coefficients. A simple parametric analysis is
performed for a typical footing resting on a halfspace soil profile and subjected to a scaled
ground motion. The dynamic impedance coefficient are estimated with an equivalent linear
procedure and compared to the linear elastic case. The dynamic stiffness coefficient is
found to decrease in amplitude and become frequency dependent, depending on the initial
soil shear wave velocity. The radiation damping is found to be unaffected by the nonlinear
soil behavior.

INTRODUCTION

normalizes the excitation frequency with a


characteristic dimension B of the foundation
and the shear wave velocity Vs of the
supporting soil.
Geometrical and material linearity is implicit
in equation (1). The static stiffness of the
footing depends on the initial shear modulus G
of the supporting soil and the Poisson's ratio ,
while the dynamic stiffness and damping
coefficients depend on the excitation frequency.
The hysteretic material damping depends on
the initial linear material properties of the soil.
Besides,
the
dimensionless
frequency
parameter 0 depends on the shear wave
velocity of the linear soil. For convenience and
to avoid any misconceptions, the shear wave
velocity of the supporting soil will be assumed
herein as the shear wave velocity Vs,30 of the
upper 30m of soil, according to the Eurocode 8.
As opposed to the idealization of the linear
soil behavior, the dynamic material properties of
the soil are not constant when the soil behaves

The importance, as well as the general


notion of the foundation dynamic impedance
function, is well established in engineering
practice, during the past three decades
(Veletsos 1971, Veletsos 1973, Gazetas 1983,
Dobry1986a, Dobry19886b, Gazetas1991a,
Gazetas 1991b, Sieffert 1992, Pecker 1997,
Mylonakis 2006). The general equation,
extracted from Gazetas 1983,
S = K ( k(,) + 0 c(,) ) (1+2 )

(1)

provides the dynamic impedance function S in


terms of the static stiffness K, the dynamic
stiffness and damping coefficients k and c
respectively, and the hysteretic material
damping ratio of an equivalent SDOF
oscillator. The dimensionless parameter
0 = B/Vs

(2)

212

in a nonlinear way. The shear modulus G and,


consequently, the shear wave velocity Vs,30
change in time with varying earthquake
excitation amplitude and frequency content.
Thus, equation (1) cannot be directly
implemented in the nonlinear soil behavior
case, since the abovementioned parameters
vary. As a result, the foundation dynamic
impedance functions have to be calculated in
terms of the modified soil-foundation system
properties. For that reason, an equivalent linear
approach can be implemented in the calculation
of the foundation dynamic impedance functions,
using equivalent linear shear modulus and
damping characteristics.
An equivalent linear procedure is used
herein to calculate the dynamic impedance
functions of a foundation supported on a soil
profile having equivalent linear properties
(Pitilakis D., 2006). The soil-foundation system
is subjected to an earthquake ground motion
consisting of body waves, propagating only on
the vertical direction. The effect of the different
initial shear wave velocity Vs,30 on the dynamic
foundation impedance functions is clearly
demonstrated.
The computation of the dynamic foundation
response is performed assuming that there is
no structure founded on the footing. Moreover,
the foundation is assumed usually massless in
a typical calculation of the dynamic impedance
functions. Therefore, no secondary nonlinear
effects are expected to appear in the soil, as
there is no diffracted wave field, created by the
foundation-structure vibration and emanating
away from the foundation to infinity. There
would still be created a diffracted wave field due
to kinematic interaction but it vanishes in the
present case for vertical incidence and surface
foundations. Assuming that the secondary
nonlinearities induced in the soil are
approximately zero, the dynamic impedance
functions estimated by the equivalent linear
procedure can be used for the computation of
the SSI with an equivalent linear soil behavior.

SYSTEM IDENTIFICATION

In a linear soil-foundation system, factors


that influence the dynamic foundation
impedances are identified by several authors.
Among them are the foundation shape and
flexibility, the embedment ratio, the excitation
frequency, the type of soil profile along with the
depth and stiffness of the bedrock, the soil
properties (shear modulus, Poisson's ratio,
hysteretic damping ratio), the soil anisotropy,
inhomogeneity and nonlinearity.
In the nonlinear soil--foundation system,
however, parameters such as the intensity and
frequency content of the signal may modify
significantly the response from one case to
another, by exciting the nonlinear response of
the soil. In contrast with the linear soil case, the
nonlinear soil behavior tends to decrease the
stiffness of the system and to increase the
energy dissipation by viscous and hysteretic
mechanisms in the soil. Thus, additional
parameters enter in the estimation of the
foundation dynamic impedance functions.
In order to reveal the effects of the nonlinear
soil behavior, and more specifically of the
different initial soil shear wave velocity, on the
foundation dynamic impedance functions, a
series of parametric analyses is performed. A
rigid
massless
footing
is
placed
a
homogeneous halfspace soil profile and is
subjected to a harmonic motion. Up to this point
the procedure is similar to the largely exploited
procedure, assuming linear soil behavior,
presented in numerous studies. The nonlinear
soil behavior is approximated by shear modulus
reduction and damping curves for a typical
clayey soil materials, notably clay with IP0.
Furthermore, the soil--foundation system is
subjected to an earthquake record and the
response is obtained in the light of the
substructure technique with equivalent linear
soil properties.
Due to the inherent complexity of such an
approach to the equivalent linear dynamic
impedance functions, several assumptions
have to be made for sake of simplicity and

213

different sizes as well as different types of


elements. More specifically, configurations were
tested using 99 quadrilateral elements, 321
quadrilateral elements, 241 triangular elements,
and the chosen configuration of 334
quadrilateral and 4 triangular elements. The
real and imaginary parts of the estimated
dynamic impedances were compared against
the reference solution proposed by Sieffert
(1992). The presented configuration (334
quadrilateral and 4 triangular elements with an
average size of 0.5m) is chosen based on the
accuracy and the time efficiency of the
achieved solution.
For a rigid, circular, massless footing resting
on a perfectly elastic halfspace, the static
stiffness of the circular disk in the horizontal,
vertical, rocking and torsional modes are
calculated according to Veletsos (1973),

comprehension.
Accordingly,
the
soilfoundation system properties are chosen so as
to cover the wider possible range of cases with
the minimum number of interfering parameters.

Foundation Identification

The foundation consists of an infinitely rigid,


massless circular footing of diameter d=10m,
resting on the free soil surface. Any arbitrary
shaped footing can be approximated by an
equivalent circular one, by equating the contact
surfaces for the three translational degrees of
freedom and the area moments of inertia for the
three rotational components.
The radius of the circular foundation, r=5m,
was chosen sufficiently large in order to
account for the kinematic interaction, along with
the inertial interaction, for a vast range of shear
wave velocities and frequencies. The circular
footing is meshed in 334 quadrilateral and 4
triangular elements with an average size of
0.5m, as shown in Fig 1.

Kx = 8Gr / (2-)
Kz = 4Gr / (1-)
K = 8Gr3 /3(1-)
Kt = 16Gr3 / 3

(3)
(4)
(5)
(6)

where G and are the initial shear modulus and


Poisson's ratio respectively for the soil,
assuming linear conditions, and r is the radius
of the footing. The static stiffness of the disk
expresses the force which is necessary to
produce a unit displacement or rotation, for the
two translational and the rotational modes of
vibration respectively.
In the vibration of a footing under loading,
the amplitude of the dynamic response in each
mode depends on the zone of influence, which
is the depth to which the normal stresses
extend in the soil. Therefore, for a circular
footing resting on a homogeneous halfspace,
the vertical normal stresses induced along the
centerline of the foundation extend to a depth of
five radii. For a horizontally loaded footing, the
horizontal stresses induced in the soil vanish at
depth larger than two radii, while for moment
and torsional loading the stresses practically
exist down to a depth of 1.25 and 0.75 radii
respectively (Gazetas 1983).

Fig 1: Rigid massless circular footing used in the


parametric analyses

Different mesh configurations were tested,


using larger or smaller number of elements,

214

naturally, in the case of a nonlinear soil


behavior and thus, an initial hysteretic damping
ratio is set at 2% for linear soil conditions. The
Poisson's ratio is set at 1/3, a typical value for
unsaturated soil.

Soil Profile Identification

The footing is placed on the free surface of


a homogeneous soil profile, as shown in Fig 2.
For convenience, the thickness of the soil
stratum is chosen equal to H=30m. For the
homogeneous halfspace shown in Fig 2, five
different shear wave velocities Vs,30 are
assigned, notably 100m/s, 180m/s, 250m/s,
350m/s, 500m/s, classifying the soil profile to
category types C and B according to the
Eurocode 8.

Fig 3: Shear modulus reduction and damping curves


for the clay IP0 soil material
Fig 2: Homogeneous halfspace soil profile
supporting the circular footing
Ground Motion Identification

In order to approximate the nonlinear


behavior of the soil with an equivalent linear
approach, a clay with plasticity index IP0,
according to Vucetic (1991), is used to describe
the clayey soil material (Fig 3). This curve is
widely used in engineering practice and
describes a typical clayey soil material.
A discretization of the soil profile in soil
layers is performed, in order to calculate the
shear strain level at different depths and vary
the shear modulus and the damping
accordingly. Therefore, the soil profile of 30m is
divided into five layers, having thickness from
top to bottom of 1m, 2m, 4m, 8m and 15m. The
soil profile in all the analyses has a typical unit
weight of 2000kg/m3. Moreover, even though in
most of the available literature on the
foundation impedance functions zero material
damping is assumed in the soil, it is confirmed
that even at small strain levels the soil behaves
in a hysteretic way (Dobry 1986a, Dobry
1986b). The latter is more pronounced,

In order to promote the nonlinear soil


behavior, the soil--foundation system is
subjected to the strong earthquake record of
Aegion, 1995 Aegion, Greece, chosen
according
to
the
European
tectonic
environment, with varying amplitude and
frequency content. The earthquake record is
scaled to maximum acceleration amplitude of
0.01g, 0.10g, 0.20g, 0.30g and 0.50g, in order
to cover a wide range of ground motion
amplitudes. However, one must keep in mind
that scaling an accelerogram might affect and
indeed
modify
its
frequency
content.
Nonetheless, in the parametric analyses
conducted, the frequency content of the input
ground motion is not as important parameter for
the nonlinear soil behavior as is the excitation
amplitude. For acceleration amplitude 0.01g,
the soil is expected to behave in a linear way.

215

An attempt is made to estimate such


dimensionless graphs for the equivalent
dynamic stiffness and dashpot coefficients.
Then, knowing the dynamic impedance
coefficients of the footing, the response of the
superstructure can be calculated from standard
rigid body dynamics.
Four components of the circular footing
vibration are examined, notably the horizontal,
vertical, rocking and torsional. The real part of
the dynamic impedance in equation (1), namely
the dynamic stiffness coefficients are noted by
kx , kz , k and kt respectively, while the
imaginary part of equation (1) expresses the
radiation damping coefficients, which are
denoted by cx, cz, cr and ct, respectively. Both
the real and imaginary parts of the impedances
are always normalized by the linear static
stiffness for the corresponding mode of
vibration, presented in equations (3) to (6). For
the real part of the dynamic impedance, the
dynamic stiffness coefficient, the dimensionless
frequency parameter, notably 0 = B/Vs,30,LIN in
a linear analysis, is enriched by a correction
factor of Vs,30,LIN/ Vs,30,EQL in order to include the
effect of the soil softening due to nonlinear
behavior. The correction factor Vs,30,LIN/ Vs,30,EQL
is indeed the ratio of the initial, linear shear
wave velocity assigned to the soil profile over
the shear wave velocity calculated for the soil
profile having equivalent linear characteristics.
Attention should be made, as will be seen, in
reading the charts, as the equivalent linear
shear wave velocity differs in each soil case,
depending on the properties of the soil and the
dynamic characteristics of the input signal.
From a practical point of view, multiplying the
dimensionless 0 parameter by Vs,30,LIN/ Vs,30,EQL
moves the dynamic stiffness coefficient
impedance curve to the right, i.e. to the higher
frequency range. Nevertheless, this horizontal
translation to the higher frequency range is the
counterpart of the shifting of the dynamic
stiffness coefficient to the lower frequency
range, due to the soil softening. The result is
the canceling of the shifting to the lower and to
the higher frequency range, that is the peak
response will appear at the same position as it
appears in the linear case.

Fig 4: Aegion, 1995 Aegion, Greece earthquake


record (PGA=4.92m/s2)

Thus, using the earthquake record with


scaled amplitudes covers sufficiently a wide
range of earthquake scenarios with scattered
amplitudes.

PARAMETRIC ANALYSES

A parametric analysis is performed in order


to investigate the effects of the nonlinear
behavior of the soil on the dynamic foundation
impedance functions. The analyses of soil-foundation systems comprise a halfspace soil
profile of a typical clay, five initial shear wave
velocities for the profile (100m/s,180m/s,
250m/s, 350m/s, 500m/s), and an earthquake
record scaled to five different amplitudes.
In the simple linear case the foundation
dynamic impedance function are affected
primarily by the parameters mentioned in a
previous section. Thus, it is immediately
understood that the complexity of the problem
increases in the nonlinear soil case as the
number of the parameters that influence the soil
response increase.
As the typical procedure for calculating the
dynamic impedances of a footing suggests
(equation (1)), the static stiffness is calculated
from equations (3) to (6) and the dynamic
stiffness and damping coefficients are
estimated from normalized graphs and charts.

216

Contrary to the dynamic stiffness coefficient


k, the imaginary part of the dynamic
impedance, namely the radiation dashpot
coefficient c times the dimensionless parameter
0, 0 c, is plotted against 0 without any
normalization. The imaginary part is plotted in
its integrity (0 c), instead of c, in order to avoid
divergence at low frequencies. For sake of
simplicity, the imaginary part of the dynamic
impedance 0 c, will be referred to herein simply
as radiation dashpot coefficient.

frequency range the magnitude of the dynamic


impedance stiffness increases with increasing
frequency, attains a peak and then decreases
following the trend of the linear case, showing a
certain
frequency
dependence.
This
contradictory behavior in the low frequency
range (dominated by the response of the
deeper soil layers to low frequency, large wave
length, pulses) is caused by an inversion of the
shear wave velocity with depth. For an
earthquake amplitude of 0.30g, the shear wave
velocity decreases with increasing depth. At a
depth of 11m, the shear wave velocity is 77m/s,
while it is 171m/s at a depth of 0.5m under the
foundation. Similarly, the shear modulus
reduces down to 10% of the maximum linear
elastic value Gmax. Thus, the low frequency
response is dominated by pulses propagating at
significantly lower shear wave velocities than
the initial linear case, resulting in a stiffness
coefficient significantly decreased in amplitude
from the linear case. On the other hand, in the
high frequency range the response is
dominated by the surface waves which are
allowed to be created. These waves propagate
in a higher shear wave velocity, with a shorter
wavelength, than the low frequency deeper
body waves, leading to a increase in the
dynamic stiffness of the footing in the higher
frequency range. In conclusion, the increase of
the dynamic stiffness coefficient with increasing
frequency is due to the decrease of the shear
wave velocity with increasing depth, after the
equivalent linear analysis. In the case of Fig 5,
the decrease of the stiffness in the high
frequency range, after the peak, is caused by
an eventual increase of the shear wave
velocity, computed to be 83m/s at a depth of
22.5m (from 77m/s at 11m depth).

Effect of Different Soil Shear Wave Velocity

The
circular
footing
rests
on
a
homogeneous halfspace. Assuming that the soil
is a clay with plasticity index IP0 and initial
shear wave velocity Vs,30,LIN=180m/s, suggests
a soil type B according to the Eurocode 8.
When the soil--foundation system is subjected
to the Aegion earthquake record the horizontal,
vertical, rocking and torsional dynamic
impedances are shown in Fig 5 to Fig 8
respectively. The impedances are plotted for
the linear case and for five different earthquake
amplitudes, as shown in the legend of the
graphs.
The horizontal dynamic stiffness coefficient
decreases from the linear case in amplitude
with increasing level of excitation amplitude (Fig
5). The fluctuations in the stiffness coefficient
curves are the apparent result of resonances
that occur in the soil. The initially homogeneous
halfspace soil profile behaves in a nonlinear
way and interfaces are formed in the soil
between layers with different impedance ratios.
Consequently, waves emanating from the
vibrating foundation are reflected on those
interfaces and propagate back towards the
footing. The result of this propagation is the
increase of the foundation motion in some
frequencies, close to the resonance frequencies
of the newly formed inhomogeneous soil. The
short and flat peaks imply no significant
impedance ratio between the formed soil layers,
whereas they appear in the resonant
frequencies of the soil.
Contrary to the linear case, in the lower

217

Fig 5: Horizontal dynamic stiffness and radiation dashpot coefficients for a homogeneous soil profile, with
clay IP0 soil, initial shear wave velocity Vs,30,LIN=180m/s, subjected to the Aegion earthquake record

Fig 6: Vertical dynamic stiffness and radiation dashpot coefficients for a homogeneous soil profile, with clay
IP0 soil, initial shear wave velocity Vs,30,LIN=180m/s, subjected to the Aegion earthquake record

218

Fig 7: Rocking dynamic stiffness and radiation dashpot coefficients for a homogeneous soil profile, with clay
IP0 soil, initial shear wave velocity Vs,30,LIN=180m/s, subjected to the Aegion earthquake record

Fig 8: Torsional dynamic stiffness and radiation dashpot coefficients for a homogeneous soil profile, with clay
IP0 soil, initial shear wave velocity Vs,30,LIN=180m/s, subjected to the Aegion earthquake record

219

The imaginary part of the dynamic


impedance, plotted in Fig 5, is the combined
effect of the radiation viscous damping and the
hysteretic material damping. Assuming a purely
elastic soil, if there were no hysteretic damping
in the equivalent linear analyses, the values of
the imaginary part of the dynamic impedance
would be zero, implying that the radiation
dashpot would be zero. On that account, the
non zero value of the radiation dashpot for zero
frequency, even for the linear case, denotes the
presence of the hysteretic damping (equal to
2% in the linear analyses). In the linear case,
the imaginary part increases at a constant rate
with increasing frequency, implying that the
radiation damping is practically independent of
the frequency. In the equivalent linear case, the
magnitude of the imaginary part of the
impedance increases from the linear case with
increasing excitation amplitude. This increase is
attributed primarily to the hysteretic material
damping of the soil, which indeed increases in
the equivalent linear soil with increasing level of
excitation. For an earthquake amplitude of
0.30g, the hysteretic damping increases up to
17% in the deeper soil layers. The radiation
viscous damping in the horizontal mode
increases with increasing frequency at the
same rate as in the linear case. This explains
the fact that the radiation dashpot coefficients
are transposed in parallel from the linear case,
apparently unaffected by the soil nonlinearity.
For the nonlinear case, the larger the excitation
amplitude, the larger the parallel transpose of
the curve to higher values.
In the low frequency range, however, the
hysteretic damping in the equivalent linear
analyses does not increase from the linear
case, causing the radiation dashpot coefficient
curves for all excitation amplitudes to be similar
in amplitude with the linear case. This arises
from the fact that in the low frequency range no
surface waves are created. The response of the
soil is dominated by the wave fields having
longer wave lengths, which create resonance
phenomena at larger depths. As in the
equivalent linear analysis the halfspace soil
profile is discretized into soil layers with refined
thicknesses closer to the surface, the increase

of the hysteretic material damping is taken into


account mainly in the upper soil layers and not
in the deeper. Thereby, the predominant low
frequency waves in the deeper soil layers do
not influence the equivalent linear soil
response, resulting in a minor increase of the
hysteretic damping from the linear case.
The effects of the soil softening due to
nonlinear behavior are more pronounced in the
vertical stiffness coefficient in Fig 6. This is
expected because of the deeper zone of
influence of the vertical normal stresses in the
soil, caused by the vertical loading of the
footing. The magnitude of kz decreases with
increasing excitation amplitude, reaching a
reduction of 80% for the case of the excitation
amplitude 0.50g. Furthermore, the nonlinear
soil is not homogeneous any more, as seen
from the peaks and valleys for values of the
dimensionless frequency parameter 0 c
Vs,30,LIN/ Vs,30,EQL less than 1. The location of the
resonant valleys coincides with the resonant
frequencies of the soil.
Concerning the vertical radiation dashpot
coefficient, the increase in the hysteretic
damping with increasing excitation amplitude, in
the medium to higher frequency range, is
apparent in the vertical mode as well. The low
frequency range is dominated by the effects of
the deeper soil layers, in which the increase in
hysteretic damping is not accounted in the
equivalent linear analysis. The radiation
damping coefficient in the vertical mode is
slightly affected by the increase of amplitude,
increasing with frequency with a bit larger rate
than in the linear case.
Similar trends are observed in the rocking
(Fig 7) and torsional (Fig 8) vibration modes
equally. The stiffness decreases from the linear
case with increasing excitation amplitude, while
it tends to become frequency independent with
increasing level of excitation amplitude. The
hysteretic damping increases with increasing
excitation amplitude while the radiation
damping is unaffected by that increase. The
radiation damping for the nonlinear cases
increases with frequency at the same rate as in
the linear case. The small, compared to the
translational modes, radiation coefficient values

220

attested by researchers for the rocking and


torsional modes, are confirmed for the nonlinear
case as well. Yet, the radiation damping
coefficients are at least 50% lower than they
are for the translational components.
Assuming a higher initial shear wave
velocity of the soil Vs,30,LIN=350m/s, the dynamic
response of the footing resting on the stiffer soil
is shown in Fig 9 for the horizontal mode, in Fig
10 for the vertical mode, in Fig 11 for the
rocking mode and in Fig 12 for the torsional
mode.
The stiffer soil causes the dynamic stiffness
coefficient decrease less from the linear case
than for the softer soil case (Vs,30,LIN=180m/s).
Evidently, the softer soil exhibits larger
deformations and the nonlinearities are more
pronounced. For the soil with Vs,30,LIN=350m/s,
the dynamic stiffness coefficients increase
constantly with increasing frequency in the
horizontal mode, in contrast with the linear
case. This is due to the inversion of the shear
wave velocity, that is the decrease of the
velocity of the body waves with increasing
depth. As opposed to the previous soil case
with Vs,30,LIN=180m/s, the dynamic stiffness
coefficient increases with frequency in the
whole frequency range, due to the constant
decrease of the shear wave velocity with depth.
The shear wave velocity decreases to 178m/s
at a depth of 22.5m. Some flat undulations
appear due to the soil inhomogeneity caused by
the nonlinear behavior and the discretization of
the soil profile. In the vertical mode, the peaks
and valleys at a dimensionless frequency 0 c
Vs,30,LIN/ Vs,30,EQL are again more pronounced
than in the horizontal mode, suggesting
stronger resonance phenomena to occur in the
soil due to the deeper zone of vertical influence.
These undulations are quite flat due to the
existence of hysteretic damping in the soil.
The radiation damping coefficients for the
horizontal and vertical modes for the stiffer soil
resemble to the ones of the softer soil, in
magnitude and in increasing rate with
increasing frequency.

For the stiffer soil the rocking and torsional


dynamic stiffnesses decrease from the linear
case less than they decrease in the softer soil
case. Nevertheless, they tend also to become
independent of frequency in the whole
frequency range. On the other hand, in the
stiffer soil the radiation dashpot coefficients for
the rocking and torsional modes attain values
less than 50% of the ones attained for the softer
soil (Vs,30,LIN=180m/s). Especially for the
torsional mode, the dashpot coefficient is
practically independent of the exciting
frequency.

CONCLUSIONS

A parametric analysis was conducted to


demonstrate the effect of the soil shear wave
velocity on the dynamic impedance functions,
when nonlinear soil behavior is accounted. The
nonlinear soil behavior was approximated by an
equivalent linear procedure and the dynamic
impedance functions of a typical footing resting
on a typical soil profile were calculated and
compared with the linear elastic soil case. The
principal findings can be summarized as
follows:
z The dynamic response of the footing for
the nonlinear case depends on more
parameters than in the linear case. The
complexity of the linear problem is
augmented in the nonlinear case by the
influence of the shear wave velocity of
the profile (it decreases in the nonlinear
case), the soil material (characterized by
the shear modulus reduction and
damping curves) and the excitation
amplitude and frequency content.
z The
dynamic stiffness coefficient
decreases from the linear case with
increasing excitation amplitude and with
decreasing initial shear wave velocity

221

Fig 9: Horizontal dynamic stiffness and radiation dashpot coefficients for a homogeneous soil profile, with
clay IP0 soil, initial shear wave velocity Vs,30,LIN=350m/s, subjected to the Aegion earthquake record

Fig 10: Vertical dynamic stiffness and radiation dashpot coefficients for a homogeneous soil profile, with clay
IP0 soil, initial shear wave velocity Vs,30,LIN=350m/s, subjected to the Aegion earthquake record

222

Fig 11: Rocking dynamic stiffness and radiation dashpot coefficients for a homogeneous soil profile, with clay
IP0 soil, initial shear wave velocity Vs,30,LIN=350m/s, subjected to the Aegion earthquake record

Fig 12: Torsional dynamic stiffness and radiation dashpot coefficients for a homogeneous soil profile, with
clay IP0 soil, initial shear wave velocity Vs,30,LIN=350m/s, subjected to the Aegion earthquake record

223

Journal of Geotechnical Engineering Division


ASCE, Vol. 112, No. 2, pp. 109-135

The dynamic stiffness coefficient for the


initially homogeneous soil profile may
increase,
decrease,
or
become
practically constant with increasing
frequency. The response of the stiffness
coefficient depends on the modification
of the soil shear wave velocity due to
the equivalent linear soil behavior. The
shear wave velocity may indeed
decrease significantly with increasing
depth, contrary to the traditional
increase, or stiffening, with increasing
depth. Therefore, an increase of the
stiffness coefficient with increasing
frequency may be as well expected,
since this inversion of the shear wave
velocity
may
be
encountered.
Nevertheless, this phenomenon is
produced only when there is a
significant decrease of the shear wave
velocity in the deeper soil layers, due to
the equivalent linear soil behavior and
its inherent deficiencies.
The dynamic stiffness coefficient
decreases with decreasing shear
modulus of the soil for a fixed level of
shearing in the soil.
In
the
nonlinear
soil
profile
inhomogeneous soil layers are formed
and,
consequently,
resonant
frequencies of the soil appear as
fluctuations of the dynamic stiffness
coefficient at those frequencies. These
undulations are of larger amplitude for
the vertical and horizontal vibration
modes, while they do not appear in the
rocking and torsional modes. The larger
fluctuations for the vertical mode are
attributed to the deeper zone of
influence of the normal stresses induced
by the response of the footing to vertical
loading.

Dobry R., Gazetas G., Stokoe II K.H. (1986)


Dynamic response of arbitrarily shaped
foundations,
Journal
of
Geotechnical
Engineering Division ASCE, Vol. 112, No. 2,
pp. 136-154
Gazetas G. (1983) Analysis of machine foundation
vibrations: state of the art. Soil Dynamics and
Earthquake Engineering, Vol. 2, No. 1, pp. 2-42
Gazetas G. (1991) Formulas and charts for
impedances of surface and embedded
foundations,
Journal
of
Geotechnical
Engineering Division -- ASCE, Vol. 117, No. 9,
pp. 1363-1381
Gazetas G. and Stokoe K.H. (1991) Free vibration
of embedded foundations: Theory versus
experiment,
Journal
of
Geotechnical
Engineering Division -- ASCE, Vol. 117, No. 9,
pp. 1382-1401
Mylonakis G., Nikolaou S., Gazetas G. (2006)
Footings under seismic loading: Analysis and
design issues with emphasis on bridge
foundations, Soil Dynamics and Earthquake
Engineering, Vol. 26, No. 9, pp. 824-853
Pecker A. (1997) Analytical formulae for the seismic
bearing capacity of shallow strip foundations in
Seismic Behavior of Ground and Geotechnical
Structures, Balkema, Rotterdam, pp. 261-268
Pitilakis D. (2006) Soil-structure interaction modeling
using equivalent linear soil behavior in the
substructure method, Ph.D. Thesis presented in
Ecole Centrale Paris, France
Sieffert J.-G., and Cevaer F. (1992) Manuel de
fonction d'impedance, Ouest Editions/AFPS,
Nantes, France
Veletsos A.S., Wei Y.T. (1971) Lateral and rocking
vibration of footings, Journal of Soils Mechanics
and Foundation Division ASCE, Vol. 97, No.
SM9, pp. 1127-1248
Veletsos A.S., Verbic B. (1973) Vibration of
viscoelastic
foundations,
Earthquake
Engineering and Structural Dynamics ASCE,
Vol. 2, pp. 87-102

REFERENCES

Vucetic M. (1994) Cyclic threshold shear strains in


soils, Journal of Geotechnical Engineering
Division ASCE, Vol. 120, No. 12, pp. 22082228

CEN (2002) Eurocode 8, Comite Europeen de


Normalisation, Prenorme ENV 1997-1
Dobry R., and Gazetas G. (1986) Dynamic
response of arbitrarily shaped foundations,

224

Fundamental Period of Sdof Systems Including Soil-Structure Interaction


and Soil Improvement
E. Kirtas1, K. Trevlopoulos1 , E. Rovithis1, K. Pitilakis1
1

Department of Civil Engineering, Aristotle University of Thessaloniki, Greece

Abstract
The effect of soil-structure interaction on the fundamental period of a structural system is
recognized by several modern seismic codes, whereas simple straightforward relationships
are provided to estimate the modified period of the building. In the present paper an effort
takes place to estimate the fundamental period of structures with surface foundation
including soil-structure interaction phenomena, utilizing 2D plane-strain numerical
simulations with FE codes. Three characteristic fixed base periods of single degree of
freedom systems are considered during the numerical investigation, while the influence of
parameters such as the soil category, superstructure mass and height values is highlighted.
Comparison with theoretical relationships and seismic code recommendations reveal the
accuracy of the employed numerical calculations, whereas a discussion takes place
regarding the observed variation of the calculated effective period values. Further
investigation regarding the effective period modification due to subsoil stiffening
interventions, highlights the necessity to consider the enhanced soil properties during the
study of the systems dynamic response.

INTRODUCTION

provide a comprehensive procedure to


incorporate interaction phenomena into seismic
design of buildings. The employed approach
involves modification of the dynamic properties
of the structure and evaluation of the response
to the prescribed free-field motion (Jennings
and Bielak, 1973). The
expected
consequences of SSI would include an increase
in the fundamental natural period of the
structure and a change (usually increase) of the
effective damping. The effective structural
period (including interaction) is calculated using
the relationship first proposed by Veletsos and
Meek (1974):

The
mechanisms
that
soil-structure
interaction (SSI) affects the dynamic response
of structures during an earthquake event have
been highlighted by several studies over the
last few decades. Soil deformation under
seismic motion is modified by the foundation
stiffness at the first stage of interaction
(kinematic part), whereas structural oscillation
imposes additional horizontal and rotational
deformations on the foundation creating
outgoing waves (inertial part), constituting
together a rather complicated phenomenon.
Depending on the foundation shape and
formation, the soil stiffness as well as the
structural dynamic characteristics, soil-structure
interaction may possess a paramount role in
the systems seismic performance. Both
induced seismic motion and structural dynamic
response may be altered during the seismic
event, changing dramatically the behaviour
especially in the case of structures with stiff
foundations on soft soil formations.
FEMA 450 regulations (BSSC, 2003)

TSSI = T 1 + k

1 + k h2
K y
K r

(1)

where T is the fixed-base period, Ky and K the


spring stiffness values to lateral and rocking
motions calculated, k and h the stiffness and
effective height of the fixed-base structure.
Seismic base shear reduction compared to the
fixed-base structure is then calculated, using an

225

appropriate expression that considers also a


seismic response coefficient based on the
effective period determined above.
Eurocode 8, Part 5 (CEN 2002) instructions
refer to the consideration of soil-structure
interaction phenomena in the cases that SSI
effects could be detrimental, such as:
a) structures where P- (2nd order) effects
play a significant role
b) structures with massive or deep-seated
foundations, such as large bridge piers,
offshore caissons, and silos
c) slender tall structures, such as towers
and chimneys
d) structures supported on very soft soils,
with average shear wave velocity Vs,max less
than 100m/s
According to EC8, dynamic soil-structure
interaction should also take into account the
non-linear soil behavior and the radiation
damping, resulting in the majority of usual
building structures to beneficial seismic
response and reduced structural bending
moments and shear forces. Nevertheless,
specific instructions on the quantitative
consideration of SSI effects are not provided in
the context of EC8.
The Greek Seismic Code (EAK 2000) on the
other hand does not mention directly soilstructure interaction issues, apart from a
requirement to take properly into account soil
compliancy at the foundation level that may
alter the dynamic response of the structure.
Yet, none of the above seismic codes
mentions the effects of soil properties
enhancement on the SSI mechanisms during
the dynamic response of the modified system.
Indeed, the fundamental period of structures,
bridge piers or other surface constructions is
expected to change in the presence of typical
soil interventions such as soil replacement and
compaction, stone columns, jet grouting,
compaction piles etc (Pitilakis et 2005, Kirtas et
al 2007).
In the first part of the present paper 2D plain
strain calculations of the effective (SSI)
fundamental period of sdof structures are
validated by comparing with corresponding
calculations determined from seismic code
recommendations or other explicit expressions
and theoretical solutions proposed in the
bibliography. In a second part the influence of
soil improvement in the modification of structure

fundamental frequency is revealed from


validated numerical simulations, whereas the
modification of the dynamic characteristics as
well as input motion alteration due to subsoil
stiffening (Fig 1) are calculated and discussed
accordingly.

Reference
System

Modified
area
G=10G

Fig 1: Application of soil stiffness enhancement


below foundation.
CALCULATION OF EFFECTIVE PERIOD VALUES

The soil-structure system employed during


the validation procedure consists of a sdof
structure with surface strip foundation (Fig 2).
The rigid foundation is considered bonded with
the soil: sliding and foundation uplift have not
been taken into consideration during the
theoretical and numerical calculations of the
SSI period. The crucial parameters that affect
significantly the intensity of the soil-structure
interaction phenomena, and subsequently the
modification of the structural dynamic
characteristics compared to the fixed-base
case, are properly accounted for during the
investigation as presented in the following
paragraphs.
Structural mass

The superstructure mass is directly related


to the development of soil-structure interaction
phenomena. Increased values of structural
mass result in enhancement of the foundation
rocking motion, due to structural oscillations
during the dynamic response of the system.
is
Normalized
structural
mass
mnorm
implemented in order to quantify the influence
of the superstructure mass to the obtained
effective structural period TSSI, according to the
expression (Wolf, 1985):
m
(2)
mnorm = str 3

226

where mstr is the superstructure mass, is the


soil density and the characteristic dimension
of the foundation, equal to half-width B of the
strip foundation employed in this paper.

for several combinations of footing geometries


and soil profiles have been recently discussed
in Mylonakis et al. (2006), focusing primarily to
bridge
foundation
cases,
whereas
dimensionless charts that provide effective
periods and damping values of soil-structure
systems have been proposed from Aviles and
Suarez (2002).
Despite the fact that numerical calculations
concern only one specific case of foundation
geometry, it is quite interesting to review the
variation
between
different
theoretical
expressions for foundation dynamic stiffness
values. Therefore, during the theoretical
calculation of stiffness parameters in the
present
study,
several
foundation-soil
combinations are examined, as presented in
Table 1 and Fig 3. The provided expressions
are adequate to estimate the effective structural
period based on Equation 1 that concern SSI
behaviour in 2 dimensions considering
horizontal and rocking motion. Details of the
approach adopted in each specific case are
summarized in the following paragraphs.

Structural height

Structural height is also strongly related to


the rocking motion of the foundation due to soilstructure interaction, increasing the oscillation
of the superstructure and enhancing effective
period modification. Normalized structural
height is considered herein to study structures
of different height, according to the expression
(Wolf, 1985):
hnorm =

hstr
B

(3)

Foundation soil compliancy

Depending on the ground type, foundation


soil compliancy may differ between different soil
categories, affecting significantly the dynamic
properties of the SSI system. In this paper three
soil types with different properties are
examined, each indicative of the soil categories
B, C and D as defined by EC8. The considered
shear wave velocities Vs are equal to 400m/s,
200m/s and 100m/s for soil types B, C and D
respectively. This correlation is useful from the
engineering point of view, during the
comparative evaluation of the obtained TSSI
values with respect to the examined soil
categories.

(a) Square foundation over uniform half-space


The expression used for the dynamic
stiffness calculation of square foundation
geometry over uniform half-space can be found
in Gazetas (1991) and Gazetas (1997). The
dynamic stiffness coefficient of the horizontal
response mode is usually considered equal to 1
for square footings, whereas the coefficient of
the rocking response mode is frequency
dependent. The dimensions of the square
footing are considered equal to the width of the
strip foundation of Fig 2.

Theoretical calculation of effective period values

Several theoretical relationships have been


proposed to estimate the soil stiffness
parameters of Equation 1. Static stiffness
values are usually first determined, and the
dynamic stiffness is then obtained by utilizing a
dynamic coefficient that is proposed for each
degree of freedom at the soil-foundation
interface. This dynamic coefficient depends on
the geometry, stiffness and embedment of the
foundation body and is usually frequency
dependent. Numerous studies in the last few
decades have investigated several cases of
foundation geometries including strip, circular,
square, rectangular or in general arbitrary
foundation shapes. Most of the studies concern
homogeneous half-space consideration of the
foundation subsoil conditions, yet layered soil
profiles over bedrock or uniform half-space
have also been studied. Analytical expressions

(b) Arbitrary mat foundation over uniform halfspace


The effective period of arbitrary mat
foundation over uniform half-space is calculated
directly according to the expression of Table 1,
proposed in FEMA 450 Commentary (BSSC,
2003). The effective weight and height of the
structure are equal to the total W and h for sdof
systems where the whole mass is concentrated
in one specific level. The value of the
foundation characteristic radius taken into
consideration during the stiffness calculations
varies, depending on the response mode as
presented in Table 1.

227

mstr
hstr
2B
Shear wave velocity VS
Shear modulus G

Density
Damping

Fig 2: Investigated soil-structure system


Table 1: Theoretical expressions for dynamic stiffness calculation
Foundationsoil
properties

Response
Mode

(a)
Square foundation
over uniform halfspace

Horizontal

Static Stiffness

Ky =

Kr =

Rocking

(b)
Arbitrary mat
foundation over
uniform half-space

Dynamic Coefficient

L
proposed
k y = k y ,0 ,

B
diagram

9GB
2v

3.6GB
1 v

effective period TSSI = T 1 +

25 r h
1.12 r h 2
1 +
2
2

Vs T
rm3

Horizontal

8Gr 1 r
Ky =
1 +
2 v 2 H

Rocking

8Gr 3 1 r
m
Kr =
1 + m
3 (1 v ) 6 H

(d)
Strip foundation on
soil layer over
bedrock

Horizontal
Rocking

2L

Gazetas 1991,
Gazetas 1997

k rx 1 0.200

(c)
Arbitrary foundation
on soil layer over
bedrock

Ky

Reference

FEMA 450
(BSSC 2003)

y 1

(Commentary FEMA 450)


H
proposed
k y = k y ,0 ,

B
diagram

2G
B
1 + 2
2v
H

Kr
GB2
B

1 + 0.2
2L 2 (1 v )
H

k r 1 0.200

FEMA 450
(BSSC 2003),
Elsabee et al. (1977),
Kausel and Roesset (1975)

Gazetas and Roesset


(1976), Gazetas (1983),
Mylonakis et al. (2006)

In the expressions of the Table:

(a) 0 =
, 2B is the footing width and G is the soil shear modulus
VS
(b) =

W
, A o is the foundation area, h and W are the height and weight of the sdof structure, dynamic rocking coefficient
A0 h

(FEMA 450), r =

A0
, rm =

4 I0
and Io is the footing moment of inertia about axis vertical to the rocking motion

(c) H is the soil layer thickness, whereas stiffness values are valid for r

(a)

< 0.5 , where for K y , r = r and for K r , r = rm

(c)

(b)

Fig 3: Foundation geometries and soil profiles examined

228

(d)

structural period. Simulation efficiency of the


codes has been previously verified in wave
propagation and site effects investigation, as
well as in selected physical centrifuge
experiments of SSI systems (Pitilakis et al.
2005, Kirtas et al. 2006a).
In order to identify the effective structural
period TSSI, the results of numerical time-history
analysis are employed. An earthquake
recording is imposed at the base of the model
(bedrock level) and seismic waves propagate
throughout the soil deposit to the surface and
the structural base, triggering therefore soilstructure interaction phenomena. The response
period of the structure in each soil-structure
combination is then identified, using the Fourier
transform of the corresponding response timehistories at the base and the top of the structure.
Indeed the top-to-base ratio of the response
histories depicted in the frequency domain
(transfer function), can be utilized for the
determination of the resonance period at the
point of maximum motion amplification. This
effective period varies for different soil types,
given the same structural fixed-base T, subject
to different soil compliances that affect the
extent of the developing interaction phenomena.
Using the Fourier ratio instead of plain
modal analysis to identify the dynamic
characteristics of the soil-structure system can
provide additional information apart from the
effective period value. An indicative example in
the case of the structure with fixed-base period
equal to 0.2s is presented in Fig 4. The
resulting effective period of the structure can be
identified quite easily from the diagram in the
case of soil categories B and C. On the other
hand, an effective period value is also
determined when soil type D is examined, by
selecting the period value at the peak of the
Fourier ratio. Nevertheless, the frequency
content of the response in this particular
combination is not very clear, since the ratio
has almost uniform value for a significant
frequency range. Additional information is
therefore provided from the Fourier ratio,
highlighting
the
enhanced
interaction
phenomena in the case of the soft soil deposit
and questioning the use of a single effective
period value to characterize uniquely the overall
structural response. This phenomenon is also
reviewed in Aviles and Suarez (2002) for
systems involving short structures.

(c) Arbitrary foundation on soil layer over


bedrock
The relationships employed to calculate
dynamic stiffness of arbitrary foundation
geometry on soil layer over bedrock can be
found in FEMA 450 (BSSC, 2003), based on
the work of Elsabee et al. (1977) and Kausel
and Roesset (1975). The same expressions
with the addition of proper coefficients can be
also employed for embedded foundations on
soil stratum over bedrock.
(d) Strip foundation on soil layer over bedrock
Dynamic stiffness in the case of strip
foundation lying on soil layer over bedrock was
calculated using the expressions that can be
found in Mylonakis et al. (2006), based on the
work of Gazetas and Roesset (1976) and
Gazetas (1983). The strip foundation on soil
stratum over bedrock investigated during the
numerical simulation is more closely related to
the specific theoretical calculation, presenting
therefore a particular interest when reviewing
the results. The dynamic coefficient for
horizontal movement presents fluctuations with
frequency, depending on the H/B ratio and the
examined frequency range. In this paper, with
the acceptance of a small error, the horizontal
dynamic coefficient is assumed equal to 1 for all
calculations. Taking into consideration the
examined H/B ratio (>4), the dynamic
coefficient values calculated here are rather
overestimated as can be observed in the
corresponding diagram in Mylonakis et al.
(2006).
Numerical calculation of effective period values

Numerical plain strain calculation of the


effective period for the structure with the strip
footing is based on the model schematically
depicted in Fig 2. The value of L/H ratio is
selected equal to 4, to avoid any undesired
reflected waves emanating from the side
boundaries, whereas H/B ratio is equal to 4.6.
Soil behavior is considered linear elastic,
characterized by the shear modulus G and
density that determine the shear wave
velocity within the soil deposit as:
G
(4)
VS =

General purposed FE codes ADINA (2005)


and ANSYS (2000) are employed for the
numerical determination of the effective

229

16

Soil type B

12

mnorm=0.5
mnorm=1
mnorm=2

Structural Transfer Function

0
0.1

0.2

0.3

0.4

12

0.5

0.6

0.7

0.8

0.9

1.0

0.7

0.8

0.9

1.0

0.7

0.8

0.9

1.0

Soil type C

0
0.1

0.2

0.3

0.4

12

0.5

0.6

Soil type D

0
0.1

0.2

0.3

0.4

0.5

0.6

T (sec)

Fig 4: Top-to-base Fourier ratios (transfer functions) for structure with T=0.2s (fixed base)
Table 2: Theoretical and numerical analysis results of effective period (hnorm=1)
Soil-mnorm case

Fixed
base
structure
T=0.2s

Theoretical

Numerical

Fixed
base
structure
T=0.4s

Theoretical

Numerical

Fixed
base
structure
T=0.6s

Theoretical

Numerical

B-0.5

B-1.0

B-2.0

C-0.5

C-1.0

C-2.0

D-0.5

D-1.0

D-2.0

(a)

0.21

0.219

0.236

0.239

0.271

0.325

0.338

0.425

0.555

(b)

0.208

0.216

0.231

0.232

0.259

0.306

0.311

0.389

0.507

(c)

0.208

0.216

0.23

0.231

0.259

0.305

0.311

0.388

0.505

(d)

0.213

0.225

0.247

0.249

0.289

0.354

0.366

0.469

0.622

ADINA

0.206

0.21

0.217

0.213

0.251

0.266

0.268

0.39

0.506

ANSYS

0.217

0.226

0.245

0.24

0.253

0.281

0.336

0.347

0.488

(a)

0.405

0.41

0.419

0.42

0.438

0.473

0.478

0.543

0.65

(b)

0.404

0.408

0.416

0.416

0.432

0.461

0.463

0.518

0.612

(c)

0.404

0.408

0.416

0.416

0.431

0.461

0.463

0.517

0.611

(d)

0.406

0.413

0.425

0.426

0.449

0.494

0.499

0.578

0.709

ADINA

0.402

0.406

0.41

0.41

0.418

0.431

0.422

0.506

0.532

ANSYS

0.406

0.41

0.422

0.422

0.44

0.471

0.482

0.506

0.546

(a)

0.603

0.606

0.613

0.613

0.625

0.65

0.652

0.7

0.786

(b)

0.603

0.605

0.611

0.611

0.622

0.643

0.643

0.683

0.757

(c)

0.603

0.605

0.611

0.611

0.621

0.642

0.643

0.683

0.755

(d)

0.604

0.608

0.617

0.617

0.633

0.665

0.668

0.729

0.835

ADINA

0.602

0.602

0.611

0.611

0.611

0.621

0.63

0.661

0.719

ANSYS

0.611

0.611

0.611

0.611

0.621

0.63

0.621

0.63

0.64

230

uniform half-space and arbitrary foundation


lying on soil layer over bedrock respectively.
Both expressions, proposed by FEMA 450
(2003), produce almost identical results,
indicating that the examined H/B ratio is rather
large to interfere to the calculation of the
effective period value. On the other hand, the
observed difference between (a) and (b) is quite
large considering that they refer to the same
subsoil conditions and equivalent foundation
properties (i.e. square and arbitrary footing
shape). Furthermore, it is evident from Table 1
that when the thickness of the soil layer
underneath the foundation is small compared to
the characteristic footing dimension, the
impedance factor is increased compared to the
case of uniform halfspace (which occurs when
H/B approaches infinity). This is not depicted in
Fig 5 where the foundation case (d) constantly
presents larger values of effective period (i.e.
smaller overall stiffness) compared to any other
case, even though the corresponding stiffness
parameters where already overestimated during
the calculations since the horizontal dynamic
coefficient was assumed equal to 1 as
mentioned previously.
It is therefore deduced from all the above
observations, that in order to make an accurate
estimation of the effective period it is not
compulsory to consider in great detail the
particular characteristics of the involved subsoil
conditions and foundation properties, at least
for cases similar to the ones examined here. A
rough description of the soil-structure system in
terms of subsoil geometry and foundation
shape, along with the proper determination of
the soil type and structural mass, height and
fixed base period, seem to provide a
satisfactory effective period value. The variation
in the case of the utilized theoretical
relationships lies in relatively small values, not
exceeding 15-20% between the different
calculations, verifying the coherence amongst
the aforementioned expressions. Taking into
consideration the inherent difficulties in the
prediction of the dynamic characteristics due to
soil-structure interaction, a more detailed
description of the systems dynamic response
than simply the estimation of TSSI may require
the thorough procedures proposed in the
corresponding references.
It is evident in all examined structural
periods that the numerical results tend to

EVALUATION OF THE RESULTS AND


VALIDATION OF NUMERICAL SIMULATIONS

Validation of effective period numerical


calculations is presented in the next few
paragraphs, based on the comparison of
numerical analyses using sdof structures on
rigid strip foundations, towards calculations that
consider theoretical expressions given in the
bibliography. The examination of the several
soil typestructural mass combinations for
each structural fixed-base period provides the
results of Table 2. In Fig 5 a comprehensive
diagram of the various investigated cases is
illustrated, presenting the variation of the
effective period for each combination. In the
same diagram both theoretical and numerical
results are presented, allowing for a
comparative evaluation of the corresponding
calculations. It is important to mention here that
increased values of structural mass or lower soil
categories are directly related to increasing
interaction effects and larger effective period
values. Nevertheless, when combinations of the
parameters soil type-structural mass are
examined (X axis in Fig 5 diagram), it is not
possible to predict in advance the case that
presents more significant period modifications.
The
sequence
of
the
soil-structure
combinations examined in the X axis of the
specific diagram is therefore indicative, since for
example C-2 case (soil type-normalized mass)
could result in larger effective period than D-0.5
combination that follows.
Natural period modification due to soilstructure interaction effects is more pronounced
in the case of lower period (stiffer) structures
with increased normalized mass. Indeed, the
structure of fixed base period equal to 0.2s
approaches effective period values of the order
of 0.5s to 0.6s, in the extreme case of very soft
subsoil conditions and large superstructure
mass that result in the development of
significant interaction effects. The modification
of the natural period in the specific structure is
equal to 150-200% compared to the fixed base
value. On the other hand, maximum effective
period alteration is reduced when more flexible
structures are examined, presenting increase of
40-75% and 17-40% in the cases of 0.4s and
0.6s structures respectively.
It is quite interesting to comment on the
calculated effective periods of cases (b) and (c)
that refer to arbitrary mat foundation over

231

Structural Period T=0.2s (fixed base)

1.00
0.90
0.80

T SSI (sec)

0.70
0.60
0.50

(a) Square on halfspace


(b) Mat found. arbitrary on halfspace
(c) Flexible arbitrary on soil layer
(d) Strip on soil layer
Numerical Analysis ADINA
Numerical Analysis ANSYS

0.40
0.30
0.20
0.10
0.00
B-0.5

B-1

0.80

T SSI (sec)

0.70
0.60
0.50

C-0.5
C-1
C-2
Soil Type - m norm

D-0.5

D-1

D-2

D-1

D-2

D-1

D-2

Structural Period T=0.4s (fixed base)

1.00
0.90

B-2

(a) Square on halfspace


(b) Mat found. arbitrary on halfspace
(c) Flexible arbitrary on soil layer
(d) Strip on soil layer
Numerical Analysis ADINA
Numerical Analysis ANSYS

0.40
0.30
0.20
0.10
0.00
B-0.5

B-1

B-2

C-0.5
C-1
C-2
Soil Type - m norm

D-0.5

Structural Period T=0.6s (fixed base)

1.00
0.90
0.80

T SSI (sec)

0.70
0.60
0.50
0.40
0.30
0.20
0.10

(a) Square on halfspace


(b) Mat found. arbitrary on halfspace
(c) Flexible arbitrary on soil layer
(d) Strip on soil layer
Numerical Analysis ADINA
Numerical Analysis ANSYS

0.00
B-0.5

B-1

B-2

C-0.5
C-1
C-2
Soil Type - m norm

D-0.5

Fig 5: Effective period variation with soil type-normalized mass combinations (hnorm=1)

232

period
modification
due
to
subsoil
interventions of limited scale in the foundation
area. Indeed several interventions result in
subsoil stiffness increase in order to enhance
soil strength and reduce settlements. The
indirect consequence of similar mitigation
methods would be the modification of the soilstructure systems dynamic properties,
altering simultaneously the input motion and
the seismic response in the case of an
earthquake event (Kirtas et al., 2006b). Yet,
as mentioned previously, seismic codes do
not require a proper consideration of the
imposed
foundation
soil
properties
modification on the input motion and the
dynamic characteristics of the soil-structure
system.
The specific intervention case examined
here, refers to a soil shear modulus increase
by a factor of 10, resulting in a much stiffer
soil formation compared to the initial
conditions at a specific area below foundation
(Fig 1). Therefore, soil category D with a
reference shear modulus value equal to 18
GPa, after the intervention is assumed to
obtain a G value equal to 180 GPa,
corresponding to shear wave velocity of 316
m/s. Such an increase implemented in soil
type B would result in a mixture that can no
longer be characterized as soil, since the
modified shear wave velocity is larger than
1000 m/s resembling properties of soft rock
formations. Thus, the theoretical predictions in
the diagrams of Fig 8 referring to soil type B
(marked as B* to denote enhanced soil
stiffness), should not be taken into
consideration since the modified soil
properties are outside the valid range of the
corresponding relationships.
Moreover, during the evaluation of the
effective period, an inherent deviation is
expected between the theoretical predictions
and the numerical simulations due to the
partial application of the soil stiffness
enhancement. Indeed, theoretical predictions
utilizing the modified soil stiffness would
actually concern a general mitigation in the
surface layer over the whole area beneath the
foundation, providing therefore erroneously
smaller effective period estimates in the
present case that only a limited area is
modified.

slightly underestimate, compared to the


existing analytical expressions, the effective
structural
period
when
soil-structure
interaction is considered. Yet, from a
preliminary evaluation of the obtained results,
both employed numerical codes seem to
follow closely the trend of the TSSI variation
with different soil-structure combinations.
Effective periods calculated with ANSYS are
closer to the theoretical solutions in the case
of 0.2s and 0.4s structures, whereas in the
case of the 0.6s structure it is ADINA
calculations that fit better the theoretical
results. In general, effective period values
calculated with ADINA appear to form a lower
bound to the examined approaches,
presenting a constant deviation compared to
theoretically calculated values. On the other
hand, ANSYS results do not present a uniform
pattern. In certain cases it coincides with the
analytical results whereas it is failing to follow
the trend of the various analytical approaches
in the specific structure of T=0.6s.
The deviation of the numerical simulation
results compared to cases (c) and (d) of the
corresponding analytical calculations is
presented in the diagrams of Fig 6. The
aforementioned conclusion regarding the
efficiency of each FE code is verified from the
specific diagram. The maximum deviation of
the numerical calculations compared to the
analytical solutions is slightly over 20%,
revealing an overall satisfactory performance
of ANSYS and ADINA on TSSI determination.
When higher structures are investigated,
the interaction effects become more
significant resulting in increased effective
period with structural height as presented in
Fig 7 for the specific structure of 0.4s fixed
base period and normalized structural mass
equal to 1. Numerical investigation with
ADINA code revealed an even better
agreement between theoretical predictions
and numerical simulations. Indeed numerical
results lie within the range of the analytical
calculations, indicating an efficient modelling
of the interaction phenomena in terms of
structural period modification.
APPLICATION OF SUBSOIL STIFFENING
INTERVENTION

An interesting question within the context


of the present paper concerns the effective

233

Numerical Calculations deviation from


foundation case (c)
100
75

Numerical Calculations deviation from


foundation case (d)
100

Tstr=0.2s (fixed-base)

50
25

25

0
-25

0
-25

-50
-75

-50

-100

-100

-75

Tstr=0.4s (fixed-base)

Deviation (%)

75
50
25
0
-25
-50
-75
-100

Tstr=0.4s (fixed-base)

75
50
25
0
-25
-50
-75
-100

Tstr=0.6s (fixed-base)

75

Tstr=0.6s (fixed-base)

75

50
25

50

0
-25

0
-25

25

-50
-75

-50

ADINA Deviation (%)


ANSYS Deviation (%)

-100

ADINA Deviation (%)


ANSYS Deviation (%)

-75
-100

B-0.5

B-1

B-2 C-0.5 C-1

C-2 D-0.5 D-1

D-2

B-0.5

B-1

B-2 C-0.5 C-1

C-2 D-0.5 D-1

Soil type - mnorm

Soil type - mnorm

Fig 6: Deviation of numerical calculations from theoretical approaches (c) (diagram on the left)
and (d) (diagram on the right)

Structural Period T=0.4s (fixed base)


1.00
(a) Square on halfspace
0.90
0.80

(b) Mat found. arbitrary on halfspace


(c) Flexible arbitrary on soil layer

0.70
(d) Strip on soil layer
T SSI (sec)

Deviation (%)

Tstr=0.2s (fixed-base)

75
50

0.60

Numerical Analysis ADINA

0.50
0.40
0.30
0.20
0.10
0.00
B-1

B-2

B-3

C-1
C-2
C-3
Soil Type - h norm

D-1

D-2

Fig 7: Effective period variation with soil type-normalized height combinations

234

D-3

D-2

Structural Period T=0.2s (fixed base)

0.60
0.55
0.50

TSSI (sec)

0.45
0.40
0.35
0.30

*(a) Square on halfspace


*(b) Mat found. arbitrary on halfspace
*(c) Flexible arbitrary on soil layer
*(d) Strip on soil layer
*Numerical Analysis ADINA
*Numerical Analysis ANSYS
Numerical Analysis ADINA
Numerical Analysis ANSYS

0.25
0.20
0.15
0.10
B*-0.5

B*-1

B*-2

C*-0.5
C*-1
C*-2
Soil Type - mnorm

D*-0.5

D*-1

D*-2

D*-1

D*-2

D*-1

D*-2

Structural Period T=0.4s (fixed base)

0.80
0.75
0.70

TSSI (sec)

0.65
0.60
0.55
0.50
0.45
0.40
0.35
0.30
B*-0.5

B*-1

B*-2

C*-0.5
C*-1
C*-2
Soil Type - mnorm

D*-0.5

Structural Period T=0.6s (fixed base)

1.00
0.95
0.90

TSSI (sec)

0.85
0.80
0.75
0.70
0.65
0.60
0.55
0.50
B*-0.5

B*-1

B*-2

C*-0.5
C*-1
C*-2
Soil Type - mnorm

D*-0.5

Fig 8: Effective period variation with soil type-normalized mass combinations (* to denote upgraded soil)
(stiffened soil in black numerical calculation of unmodified soil cases are in grey colour)

235

constantly underestimate the effective period


values, due to the inherent consideration of an
overall application of the soil properties
enhancement. Therefore, in similar intervention
cases, numerical investigation is deemed more
appropriate even for a rough estimation of the
TSSI values.

Observation of the effective period variation in


Fig 8 verifies the above prediction. Theoretical
relationships result systematically in smaller
effective period values compared to the
numerically obtained results. The inadequacy of
the analytical relationships to predict the TSSI
value without the proper consideration of the
limited intervention extent is therefore evident,
especially when soft soil conditions and large
superstructure mass characterize the initial soilstructure case. The relative deviation between
ADINA and ANSYS is similar to the one
observed in the reference soil-structure
systems, yet this time forming the upper bound
of the illustrated values. Comparison of
numerical results (grey and black lines in Fig 8)
is indicative of the significant intervention effect,
leading the structural response to frequencies
closer to the fundamental frequency of the fixed
base structure.

REFERENCES
ADINA (2005) "Automatic Dynamic Incremental
Nonlinear Analysis. Theory and Modelling Guide",
ADINA R&D, Inc.
ANSYS (2000) "ANSYS Users manual. Version 8.1",
SAS IP Inc., Houston, USA
Aviles J., Suarez M. (2002) "Effective Periods and
Dampings of Building-Foundation Systems
Including Seismic Wave Effects", Engineering
Structures, Vol. 24, 553-562
Building Seismic Safety Council (2003), "FEMA 450
- NHRP Recommended Provisions for Seismic
Regulations for New Buildings and Other
Structures", Federal Emergency Management
Agency, Washington D.C.

CONCLUSIONS

The effective structural period in several


soil-structure cases has been estimated using
well-known available theoretical expressions
and numerical simulations with FE codes
ADINA and ANSYS. Subsoil conditions refer to
soil categories B, C and D according to EC8,
whereas three structural fixed base periods as
well as three different superstructure masses
have been implemented during the investigation
to include a wide range of representative soilstructure cases. Validation of the effective
period calculations from the numerical
simulations that was based on theoretical
solutions, revealed a satisfactory approximation
of the structural period modification due to soilstructure interaction. Application of the same
comparison in soil-structure cases of locally
mitigated subsoil area beneath the foundation,
indicates the induced reduction of the effective
period due to similar interventions of enhanced
soil stiffness. Numerical investigation revealed
that the imposed alteration of the systems
dynamic characteristics, due to the employed
mitigation, may result in a significant
modification of the seismic response, especially
in cases of stiff structures founded on soft soil
conditions. On the other hand, the inadequacy
of the analytical relationships to predict the TSSI
value without the proper consideration of the
limited
intervention
extent
is
evident.
Calculations based on analytical relationships

CEN (2002) prEN 19985: Eurocode 8: Design of


Structures for Earthquake Resistance, Part 5:
Foundations,
Retaining
Structures
and
Geotechnical Aspects, European Committee for
Standardisation, Brussels
Elsabee F., Kausel I. and Roesset J.M. (1977)
"Dynamic Stiffness of Embedded Foundations",
Proceedings of the ASCE Second Annual
Engineering Mechanics Division Specialty
Conference, pp. 40-43
Gazetas G. (1983) "Analysis of Machine Foundation
Vibrations: State of the Art", Soil Dynamics and
Earthquake Engineering, Vol. 2, No. 1, pp. 2-42
Gazetas G. (1997) "Dynamic Soil-Structure
Interaction During Earthquakes", Proceedings of
the Advanced Study Course on Seismic Risk
(SERINA), ITSAK, Thessaloniki, Greece
Gazetas G. (1991) "Formulas and Charts for
Impedances of Surface and Embedded
Foundations",
Journal
of
Geotechnical
Engineering Division, ASCE, Vol. 117, No. 9, pp.
1363-1381
Gazetas G. and Roesset JM. (1976) "Forced
vibrations of strip footings on layered soils",
Proceedings of the Specialty Conference on
Methods of Structural Analysis, Dynamic SoilStructure Interaction Session, ASCE, Vol. 1, pp.
115-131

236

Jennings P.C., Bielak J. (1973) "Dynamics of


Building-Soil Interaction", Bulletin of the
Seismological Society of America, Vol. 63, No. 1,
pp. 9-48
Kausel E. and Roesset JM. (1975) "Dynamic
Stiffness of Circular Foundations", Journal of the
Engineering Mechanics Devision, ASCE, Vol.
101, No. 6, pp. 771-785
Kirtas E., Rovithis E., Pitilakis K. and Sextos A.
(2006a) "Numerical Investigation of Potential
Foundation Intervention as a Means for
Mitigating Seismic Risk", Proceedings of the 8th
U.S. National Conference on Earthquake
Engineering, San Francisco, California
Kirtas E., Rovithis E. and Pitilakis K. (2006b)
"Numerical Investigation of Subsoil Interventions
Towards Structural Seismic Risk Mitigation",
Proceedings of the First European Conference
on Earthquake Engineering and Seismology,
Geneva, Switzerland
Kirtas E., Trevlopoulos K., Rovithis E., Pitilakis K.
(2007), "Discussion on the Fundamental Period
of Sdof Systems Including Soil-Structure
Interaction", Accepted for presentation in the 4th
International
Conference
on
Earthquake
Geotechnical Engineering, Thessaloniki, Greece
Ministry of Public Works (2000) "Greek Seismic
Code, EAK 2000", Athens (in Greek)
Mylonakis G., Nikolaou S. and Gazetas G. (2006)
"Footings Under Seismic Loading: Analysis and
Design Issues with Emphasis on Bridge
Foundations", Soil Dynamics and Earthquake
Engineering, Vol. 26, No. 9, pp. 824-853
Pitilakis K., Kirtas E. and Rovithis E. (2005) "Is it
Possible to Improve the Seismic Structural
Behaviour with Intervention to Subsoil and
Foundation Conditions?" Proceedings of the 1st
Greece-Japan Workshop: Seismic Design,
Observation and Retrofit of Foundations, 185202, Athens, Greece
Veletsos AS, Meek JW. (1974) "Dynamic Behaviour
of Building-Foundation Systems", Earthquake
Engineering and Structural Dynamics, Vol. 3, No.
2, pp. 121-138
Wolf JP. (1985) Dynamic soil-structure interaction,
Prentice-Hall, Englewood Cliffs, NJ

237

Analytical Study on Seismic Design of Footings


Subjected to Earthquake Loading
K. Kosa1, T. Ando2, Y. Adachi3, M. Shirato4
1

Kyushu Institute of Technology, Japan


2
Hanshin Expressway Co., Japan
3
Hanshin Expressway Co., Japan
4
Public Works Research Institute, Japan

Abstract
The upper part of the footing on the loading side suffers tension damage when a bridge pier
is subjected to earthquake loading. To find the mechanism of this damage, a seismic
loading test was conducted and the obtained results were evaluated using two-dimensional
elasto-plastic finite element method (FEM) analysis. It was found that the upper part of the
footing on the loading side is damaged by the bending-preceding type shear, and that this
damage occurs after the reinforcement in the upper footing has reached yielding due to a
horizontal shear force. It was also found that this shear damage usually does not result in a
brittle fracture owing to the confinement effect provided by the main reinforcement in the
lower part of the footing.

INTRODUCTION

conducted a seismic behavior simulation test


and evaluated the obtained results using twodimensional elasto-plastic FEM analysis.

When a bridge pier with a pile foundation is


subjected to earthquake loading, both the upper
and lower parts of the footing (hereinafter
referred to as upper and lower footings) on the
loading side are tensioned due to an inertia
force occurring on the pier. The behavior of the
lower footing under this loading has been
studied well1). On the other hand, the behavior
of the upper footing which should be designed
against tensile force from the column and piles,
has been experimented little and its damage
mechanism is still largely unclarified.
After the Hyogo-ken Nambu Earthquake
(Mw=6.9) in 1995, the horizontal seismic
coefficient in the Japanese Specification for
Highway Bridges was increased to twice the
conventional coefficient. Several studies also
reported that the footing shape is predominantly
determined by the resistance from the upper
footing on the loading side2). Therefore, for the
effective seismic design of footings, it is
important
to
understand
the
damage
mechanism of the upper footing under a tensile
force. As one such effort, the authors

EXPERIMENTAL PROGRAM

Specimens and experimental method


Specimens were constructed to be a
monolithic type having a pier, a footing, and
piles. They were constructed to 1/3 and 1/2
scales of an actual bridge which sustained C
rank damage (slight damage) in the Hyogo-ken
Nambu Earthquake. Fig. 1 shows the structure
of specimens. Table 1 shows their attributes.
The effective depth of footing was the same for
the specimens of the same scale, but their
shear span ratio (a/d) was changed to 0.39,
0.75, or 1.0. The reinforcement ratio on the
upper and lower footings of Specimen No. 3
was increased to 1.5 times that of Specimen No.
1. Load was applied by two axial loading
manner. Namely, a horizontal load equivalent to
a seismic force was applied to the upper part of
the column, while a vertical load of 1.6 N/mm2
equivalent to the reaction force of the

238

Table 1: Attributes of specimen


PC steel bar

No.1

Vertical jack

a/d
Width (mm)
Depth (mm)
Footing Reinforce Upper
ment ratio
Lower
(%)
Spec. for
reinforcement

200

Horizontal jack
Reaction
wall

Spesimen

a=300

340

640

Loading
direction

2000
3150

550

0.266

0.320

0.75
1830

1990 Spec.

No.3
0.75
1830
0.120

No.4
1/2
0.39
2500
750
0.075

0.408
0.144
1.5 x 1990
1990 Spec.
Spec.

a=300

340

500 450

640

0.080

No.2
1/3
1.00
2030
450
0.108

Scale

340

Unit:mm

300 550 300


340
1150
1830

Upper face of
specimen

Fig. 1: Structure of specimen

Fig. 2: Damage to the upper footing

Damage to the footing


1) Damage to the upper footing
In each specimen, dominant damage
occurred in the upper footing on the loading
side, as shown in Fig. 2. The damage
developed in the following manner.
Cracking started at the intersection of the
upper footing and the column bottom. They
expanded to the right and left sides and to the
load application side.
The main reinforcement in the upper footing
sustained a yield strain at the foot of the column
on the loading side. Almost concurrently, the
main reinforcement in the lower footing on the
compression side also sustained a yield strain.
The load reached its maximum and cracks
at the corner of the column bottom started to
propagate towards the corner of the footing.
After reaching the maximum load, the loaddisplacement curve began to decrease
gradually. Then, the upper footing on the
loading side began to rise, and a corn-shaped
failure linking the corner of the column bottom
and the corner of the footing was clearly found
(the black area in Fig. 2).

Horizontal load (kN)

superstructure was being applied to the column


top as a uniform load.

Fig. 3: Horizontal load-displacement


relationship at the loading position

relationship at the horizontal load application


position. After the main reinforcement in the
upper footing yielded and the maximum
strength is reached, the load began to decrease
gradually. In the case of Specimen No. 2 with a
large shear span ratio, the behavior was similar
up to the maximum strength, but a punching
shear failure occurred in the footing during the
decrease of load.

2) Load-displacement relationship at the


horizontal load application position
Fig. 3 shows the load-displacement

239

3) Damage to the cross section of footing


A core was taken from the cross section of
the footing in Specimen No. 4. Fig. 4 shows the
damage sustained. A diagonal crack due to the
tensile force from the column and piles is seen.
As the upper footing on the loading side rose,
this tensile shear crack appears to be a
dominant failure cause of this specimen.

Main reinforcement in the column

Column

Shear damage due to


tensile force

4) Footing damage vs. ductility


Table 2 shows the strength and ductility of
each specimen. From the damage observed, it
is considered that the load decrease was
caused by the tensile shear damage of the
footing on the loading side. If the strength
reduction up to yielding of main reinforcement
in the upper footing is taken as the ductility of
the footing, a ductility of about 2 can be
expected.

Pile

Fig. 4: Damage to the footing (No. 4)


Table 2: Strength and ductility of specimens
Hor. Load
(kN)
Yield strength
Hor. displacement
(mm)
Hor. load
(kN)
Maximum
strength
Hor. displacement
(mm)
Hor. load
(kN)
100% x yield
strength
Hor. displacement
(mm)
max/y
Ductility
100%/y

ANALYTICAL PROGRAM

Analysis method
For the analysis, two-dimensional elastoplastic FEM analysis was used. The width of the
footing was made to b (width of column) + d
(effective depth of footing) = 940 mm from the
range where the main reinforcement in the
upper footing reached yielding4).
As the loading conditions, a monotonic
load was applied to the upper side of the
column by the displacement control method
while a load equivalent to the dead weight of
the superstructure was being applied to the
column top. As the boundary condition, the
column bottom was completely fixed.

No.1

No.2

No.3

No.4

Py

245

338

342

814

22

21

25

25

Pmax

273

370

367

1032

max

38

33

40

60

P100%

245

338

342

814

100%

54

45

60

150

1.7

2.1

1.6

2.4

2.5

2.3

2.4

6.0

Table 3: Characteristics of concrete and


reinforcement
Compressive
strength
(N/mm2)
28.028

Column
Footing, pile

Material model
As the elements, plane stress elements
and linear elements were used for the concrete
and
reinforcement,
respectively.
The
reinforcement and concrete were assumed as
completely bonded. Table 3 shows the
materials characteristics of the concrete and
reinforcement. As the failure criteria, the
Drucker-Prager criterion was used for the
compressive side of the concrete and the
maximum principal stress criterion for the
tensile side5). Fig. 5 shows the stress strain
model of the concrete5). In the compression
increase range, a quadratic curve was used up
to the compressive strength, and after this
strength the stress was made to decrease

Reinforcement

Tensile
strength
(N/mm2)
2.107

Modulus of
elasticity
(N/mm2)
1.35104

25.284
1.637
1.24104
Modulus of
Yield strength Yield strain
elasticity
()
(N/mm2)
(N/mm2)
5
345
1640
2.010

Poissons
ratio
0.2
0.2
Tensile
strength
(N/mm2)
490

ft
cu

4cu

t
cr

= f c(

c
0 . 002

)( 2

c
0 . 002

0.75GF/(ftL)
5GF/(ftL)

0.2fc
)

fc
Fig. 5: Stress-strain of concrete

240


u
y

Es
100
Es

Fig. 6: Shear reduction coefficient


Fig. 7: Stress-strain of reinforcement

linearly. In the tensile range, the stress was


assumed to increase linearly up to the
maximum tensile stress (ft), and after that a 1/4
model which took softening and failure energy
into account was used. The failure energy of
the concrete Gf was assumed as 0.1N/mm. The
diagonal line of an element was used as the
equivalent length (leq) of an element which is
required for analysis. Fig. 6 shows the
relationship between the shear stress transfer
coefficient and the strain after the occurrence of
cracks.6) As shown in Fig. 7, the model for the
stress-strain relationship of reinforcement had a
yield plateau range of 8.5 times the yield strain
after reaching the reinforcement yielding, and
then took strain hardening into account.

Fig. 8: Load-displacement relationship at the


horizontal load application position

ANALYTICAL RESULTS AND DISCUSSIONS


In the analysis, all specimens showed a
behavior similar to an experimental behavior
until around the maximum load when damage
occurred to the footing on the loading side.
Taking Specimen No. 1 as the representative,
this process is reviewed in detail. Here, the
tensile strain is given as positive and the
compressive strain as negative.
Behavior up to flexural yielding
1) Horizontal load-displacement relationship
Fig. 8 shows the load-displacement relationship
at the horizontal load application position. In the
experiment, the horizontal displacement
increased after the reinforcement in the upper
footing yielded, and the load began to decrease
after reaching the maximum strength. In the
analysis, the horizontal displacement also
increased after the yielding of reinforcement,
but the load decrease was not reproduced due

Fig. 9: Distribution of maximum principal strain in the


footing on the loading side

241

Column
Upper
reinforcement
Assumed plane for
shear damage

4000
Lower
reinforcement

2000

Maximum principal strain ()

Fig. 11: Distribution of strain in the upper footing on


the loading side

Fig. 10: Cracks under the maximum load

1000
0

Pile

(When : 14 mm displacement)

Fig. 13: Distribution of maximum principal strain in


the footing on the loading side

Fig. 12: Distribution of maximum principal strain in


the footing on the loading side

3) Strain of main reinforcement in the upper


footing on the loading side
to divergence. This is probably due to a small
amount of reinforcement in the upper footing
and the effect of two-dimensional modeling.

Fig. 11 shows the distribution of strain in the


main reinforcement in the upper footing. Good
agreement was seen between the analytical
and experimental results concerning the
distribution of strains and the occurrence of a
large strain at the outer rim of the column
bottom on the loading side as the load
increased.

2) Maximum principal strain in the upper


footing on the loading side
Fig. 9 shows the maximum principal strain
in the upper footing on the loading side when
the displacement was 9 mm. Fig. 10 describes
the distribution of cracks under the maximum
load in the experiment. In the experiment, the
dominant cracks occurred in the upper footing
and flexural cracking started at the front side of
the column. In the analysis, a maximum
principal strain in the horizontal direction also
occurred in the upper footing on the loading
side. It is also presumed that flexural cracks
occurred at the front side of the column.

Behavior after flexural yielding


1) Shear damage to the footing on the
loading side
Fig. 12 shows the distribution of maximum
principal strain in the footing on the loading side
when the displacement was 14 mm. In the
experiment, diagonal cracking penetrating from
the upper face to the lower face of the footing

242

Horizontal force in the column


bottom-pile head range
Horizontal shear force
Horizontal force at the pile head
Horizontal force at the pile head

Strain in the lower reinforcement ()

400
300
200
100
0
-100
0

Fig. 14: Distribution of main reinforcement in the


lower footing on the loading side

Horizontal shear force

5
10
15
Displacement at the loading position (mm)

20

Fig. 16: Relationship between horizontal shear force


and loading position

Horizontal force at
the column bottom

Vertical force by
column reinforcement

Vertical force by
column reinforcement

Horizontal force at
the pile head

Cross section for check


of horizontal shear force

Vertical force
at the pile head

Fig. 15: Schematic of shear forces acting on the footing

was confirmed, as seen in Fig. 10. In the


analysis, a maximum principal strain in the
diagonal direction which was different from the
flexural cracking was caused at the shear span
position indicated by the circle in the figure.
Because this crack was considered as a shear
crack from its extension in the diagonal
direction, a shear damage cross section which
was perpendicular to the strain was assumed.
Fig. 13 shows the relationship between the
maximum principal strain occurring on the
assumed shear damage cross section and the
displacement at the loading position. The
maximum principal strain here is the mean
value of principal strains which occurred in the
element crossing the assumed shear damage
cross section. Strain reaching the maximum
tensile stress did not occur until yielding of

upper reinforcement when the footing was


considered to be taking flexural behavior. But,
after the yielding of upper reinforcement, strain
increased significantly.
Fig. 14 shows the distribution of strain in
the reinforcement in the lower footing on the
loading side. While the load was small,
compressive
strain
occurred
in
the
reinforcement in the lower footing. But, after the
yielding of reinforcement in the upper footing,
tensile strain also occurred in the reinforcement
in the lower footing in both the experiment and
analysis.
2) Shear force acting on the footing
Fig. 15 shows a schematic of the shear
forces acting on the footing. Dominating forces
contributing to the shear damage of the footing
were the shear force in the horizontal direction

243

Horizontal shear force (kN)

due to acting inertia force and the shear force in


the vertical direction due to pullout action of the
column and piles. From the fact that the shear
damage shown in Fig. 12 occurred after the
yielding of reinforcement in the upper footing,
and then the tension force also occurred in the
reinforcement in the lower footing, it is
considered that the force in the direction the
main reinforcement is affected, namely, the
shear force in the horizontal direction, is
predominant.
3) Horizontal shear force acting on the
footing on the loading side
Fig. 16 shows the relationship between the
horizontal shear force and displacement at the
loading position. Here, the acting horizontal
shear force was calculated by subtracting the
horizontal force acting on the pile head from the
horizontal force acting on the column bottom.
With the increase of horizontal displacement,
acting horizontal shear force increased
monotonically without showing an abrupt
increase or decrease.
Fig. 17 shows the relationship between the
ratio of horizontal shear force carried by the
concrete
and
reinforcement
and
the
displacement at the loading position. The shear
force carried by the reinforcement was
calculated from the stress acting on the
reinforcement which crosses the cross section
for the horizontal shear force check shown in
Fig. 15. The shear force carried by the concrete
was obtained by subtracting the shear force
carried by the main reinforcement from the
acting shear force.
In the elastic state, most of the horizontal
force was carried by the concrete. But, this ratio
began to decrease from around a displacement
of 7 mm when cracking appeared on the upper
footing, and instead the ratio carried by the
reinforcement
increased.
Therefore,
the
behavior in this range is considered to be close
to a flexural behavior.
In the case of general flexural behavior, a
major compression force acts on the
reinforcement in the lower footing after the
reinforcement in the upper footing has reached
yielding. But, here a tension force was
generated, and the ratio of shear force carried
by the reinforcement in the lower footing
increased and the ratio of shear force carried by

Fig. 17: Horizontal shear force carried

Fig. 18: Schematic of shear damage to the footing

the concrete decreased. At this time, shear


damage was probably caused in the footing.
After this damage, the main reinforcement in
the lower footing resisted the horizontal shear
force.
4) Mechanism of shear damage to the
footing
Fig. 18 shows a schematic of shear damage
to the footing due to a horizontal shear force. It
is considered that shear damage to the footing
is caused by the horizontal shear force acting
on the column and piles. When this force acts,
the upper footing becomes tensioned due to the
rotational behavior of the column. But, after the
reinforcement in the upper footing has yielded,
only the concrete must resist the shear force
and soon suffers shear damage at the middle of
the footing. Even though shear damage has

244

REFERENCES

occurred, a brittle fracture will not be caused


because the main reinforcement in the lower
footing still resists.

1) Shirato, M., et al. (2001) Study on the Shear


Strength Equation for Deep Beam-Footing
Connection,
Journal
of
Structural
Engineering (Japan), Vol. 47A, pp. 13151325
2) Fujii, Y., et al. (1999) Analysis of Shear
Effect on the Footing under New Highway
Specifications, Proc. of JCI, Vol. 21, pp.
1231-1236
3) Kosa, K., et al. (2000) Experimental Study
on the Behavior of Footing under Seismic
Loading with a Focus on the Shear Span
Ratio, Journal of Structural Engineering
(Japan), Vol. 46A, pp 1405-1412
4) Road Association of Japan (1996) Specifications
for
Highway
Bridges,
IV
Substructure
5) Chen, W.F. (translated by Shikibe, Kawazumi,
and Adachi) (1985) Plastic Analysis of
Concrete Structures
6) Rots, J.G. (1988) Computational Modeling of
Concrete Fractures Dissertation, Delft
University of Technology

CONCLUSIONS
The following conclusions were drawn from
the experiment and two-dimensional plastoelastic FEM analysis performed on the footing
subjected to earthquake loading.
(1) Shear damage to the upper footing on the
loading side is often preceded by the flexural
damage, because the ratio of the main
reinforcement in the upper footing is small.
(2) Shear damage to the footing occurs after
the main reinforcement in the upper footing
has reached yielding.
(3) Damage due to horizontal force does not
result in a brittle fracture owing to the
confinement effect provided by the main
reinforcement in the lower footing.

245

Development of the sheet-pile foundation;


a series of seismic proving tests
S. Higuchi1, H. Nishioka2, K. Tanaka1, M. Koda2, J. Hirao3
1

Technical Research Institute of OBAYASHI Co., Japan


2
Railway Technical Research Institute, Japan
3
Civil Engineering Technology Dept., OBAYASHI Co., Japan

Abstract
ABSTRACT: The authors have proposed the sheet-pile foundation as a new foundation type.
This foundation has advantages as follows; 1) Wider applicability to various soil conditions
than the shallow foundations, 2) Limited impact to the environment with construction work,
3) More economical than the pile foundations in terms of cost. In this paper, investigations
of the fundamental characteristics of the sheet-pile foundation during the earthquake for the
development of the design method of this structure are presented. A series of static loading
tests (laboratory and full-scale in the field) and centrifuge tests were carried out. Results
show the sheet-pile foundation has excellent performance against seismic force.

INTRODUCTION

Recently, development of construction


methods for densely populated urban area is
emphasized in Japan. For example, in order to
ease traffic congestion, railroads are re-laid on
viaducts. For this project, structures are usually
constructed very close to existing structures,
and the space allowed for construction work is
limited. In addition, it is required to reduce costs,
as well as minimizing the impact to the
environment, such as noise, vibration and
disposals from construction work.
Sheet-pile Foundation (SPF, hereafter),
which combines the footing and sheet-piles,
proposed as a new foundation form (Koda et al.
2003, Nishioka et al. 2004) is one solution.
Because of the confinement of the ground is
increased by the sheet-piles, both bearing
capacity and horizontal resistance of the SPF
are improved compared to those of the shallow
foundation. Therefore, the applicability became
wider than that of the shallow foundations. For
example, SPF can be adopted on the loose
sandy ground to which the pile foundation has
been usually applied. The construction cost of
SPF is almost the same as that of the shallow
foundation and more competitive than that of

246

the pile foundation. On the other hand, since


the pile work is not necessary, it can avoid
various disadvantages of pile foundation, such
as noise, vibration and the disposal of surplus
soil. Fig 1 shows an outline of the sheet-pile
foundation com-pared with the shallow
foundation and the pile foundation.
In this paper, a series of static loading tests
(laboratory and full-scale in the field) and
centrifuge tests carried out for the purpose to
evaluate the performance of the sheet-pile
foundation are presented and discussed.
Shallow
foundation

Sheet-Pile
foundation

Pile
foundation

Sheet-Piles

Dense Sand
(SPT N-value >30)

Loose Sand
(10< N-value >30)

Fig 1: Outline of the sheet-pile foundation.

HORIZONTAL LOADING TESTS

block of 100 mm in width B, and placed on the


model ground surface. Model sheet-pile was
made of phosphor bronze plates with 0.2 mm
thickness, and they were pressed to concavoconvex form shown in Fig 2. The length L that
was installed into the ground was 100 mm
(L/B=1.0) or 50 mm (L/B=0.5). The value of L
shown in Eq.1 of the model sheet-piles was the
same grade as that of prototype.
(1)
L = 4 k h D 4 EI L

1) Outline of model ground and foundations


A model ground was prepared in a rigid
container with dry sand. The model was a twodimensional in plane strain condition. The sand
containers side walls were made of transparent
acrylic plates to allow observation of the
deformation of the ground. In order to reduce
friction between the acrylic plate and sand,
rubber membranes were pasted on acrylic plate
with grease. Target points were marked on the
rubber membrane for measuring displacement
of the ground by an image processing system.
Table 1 summarizes the conditions of the
model ground. The relative density Dr of the
model ground was controlled by the height of
the sand hopper to 90% or 60%.
The model footing was made of aluminum

where : Characteristic value of pile (1/m), kh:


Coefficient of horizontal subgrade reaction
(kN/m3), D: Width of sheet-piles (m), EI:
Flexural rigidity of sheet-pile (kNm2), L: Length
of sheet-pile (m). Table 2 summarizes
specifications of the model sheet-piles.
2) Test procedures
In order to simulate the effects of inertia
force by the earthquake, the horizontal loading
tests were conducted. The horizontal loads
were applied to the bridge pier top.
An outline of the tests is shown in Fig 3.
The horizontal displacement was applied with a
screw jack statically and cyclically at a height of
230mm from the footing model bottom that was
corresponding to the bridge pier top. The
vertical load was applied to the pier top at 1.2
kN by an air cylinder, which was about 10% of
the bearing capacity of the shallow foundation
previously tested on the dense ground model.

Table 1: Conditions of modeled ground


Ground size
(W H D)
Material of ground
Dry unit weight d
Lubricated layer

2000 mm 580 mm
600 mm
Dry Toyoura sand
d = 16.2 kN/m3 (Dr = 90%)
d = 15.1 kN/m3 (Dr = 60%)
Rubber membrane (t=0.2mm)
with Grease (10m)

Horizontal
Displacement

Vertical Load
V = 1.2 kN

Fig 2: Picture of model sheet-pile


Table 2: Specifications of model sheet-piles

Thickness
Height of concavoconvex form
Width of Footing B
Width of
sheet-piles D
Young's modulus
E
Geometrical moment
of inertia I
Coefficient of
horizontal subgrade
reaction kh
Length L
L

Prototype
TYPE IV

Modeled
sheet-pile

15.5 mm

0.2 mm

340 mm

1.5 mm

4.8 m

100 mm

4.8 m
200 kN/mm
( Steel )

B=100mm

596 mm
2

1.2810-3 m4

D = 596 mm

h=230mm

M odel ground
Dry Toyoura sand

L=100mm
2

110 kN/mm
( Phosphor bronze )

Fig 3: Outline of horizontal reciprocal loading test.

42.9 mm4

78,600 kN/m3
( Sand N = 30)

45,700 kN/m3
( Dr=90% )

4.8 m (L/B=1.0)
2.4 m (L/B=0.5)
3.74 (L/B=1.0)

100 mm (L/B=1.0)
50 mm (L/B=0.5)
3.42 (L/B=1.0)

Sheet-Pile model

Table 3: Cases of horizontal loading tests

247

Case

Density of ground

Foundation Form

HD1
HL1
HL2
HL3

Dr=90% (Dense)

Shallow foundation
Shallow foundation
Sheet-pile foundation L/B=0.5
Sheet-pile foundation L/B=1.0

Dr =60% (Medium
dense)

0.4

of more than 10% of footing width, the SPF on


the same ground model (Case-HL2, HL3) has
only small settlements, which are as same as
that of shallow foundation on the dense ground
(Case-HD1). Therefore, it is clear that the
sheet-piles restrained the settlement.
Photo 1 shows the deformation of the
ground in two cases (Case-HL1 and HL2),
computed by the image processing system
previously described. The lines in the figures
show the locus of each target point until the
horizontal displacement at the top of pier
reaches 20mm. The deformation of the ground
was observed in a large area around the model
footing. In the case of the shallow foundation
(Case-HL1), the ground failed like a circular slip
Displacement (mm)
0

case-HL1 (SF)
case-HL3 (SPF L/B=1.0)

0.3

10

15

20

2
4
6
8
10

0.2

Load (kN)

Settlement (mm)

3) Test results
The relations between the horizontal load P
and displacement at the top of pier are shown
in Fig 4 and Fig 5. Fig 4 shows hysteresis
curves, and Fig 5 shows skeleton curves, which
connected the turning points on each loading
cycles together. The SPF (Case-HL2, HL3) has
higher horizontal resistance than that of shallow
foundations (Case-HD1, HL1). Since the loop of
the hysteric curve of the SPF is larger than that
of the shallow foundation, it is clear that the
hysteric damping of the SPF is larger than
shallow foundations. In addition, the residual
horizontal displacement of the SPF was almost
negligible after the experiment.
The settlement characteristic is another
important function of the railway structure. Fig
6 shows settlements of footings when the
horizontal displacement reached to the peak in
each cycle. Settlements became larger as the
increase of horizontal displacement in all cases.
Although the shallow foundation on the medium
dense ground (Case-HL1) had large settlement

case-HD1 (SF)
case-HL1 (SF)
case-HL2 (SPF L/B=0.5)
case-HL3 (SPF L/B=1.0)

12

Fig 6: Skeleton curves of settlements

0.1
0.0

-0.1
-0.2
-0.3
-25

-20

-15

-10

-5

10

15

20

25

Displacement (mm)

Fig 4: Hysteresis curves of P- relationship


a) Case-HL1: Shallow foundation

0.35
0.30

Load (kN)

0.25
0.20
0.15
case-HD1 (SF)
case-HL1 (SF)
case-HL2 (SPF L/B=0.5)
case-HL3 (SPF L/B=1.0)

0.10
0.05
0.00
0

10

15

20

b) Case-HL2: Sheet-pile foundation (L/B=1.0)

Displacement (mm)

Photo 1: Displacement computed by the Image


Processing System when horizontal displacement
reaches 20mm. (Dr=60%)

Fig 5: Skeleton curves of P- relationship

248

8. Sheet-pile length and ground density are


chosen as parameters of this experiment. Table
4 summarizes the conditions of the model and
ground. Shake table tests were carried out
utilizing the input motions as shown in Table 5.
2) Test results
Distribution of the vertical self load between
the sheet-piles and the footing bottom after the
centrifuge acceleration process are summarized
in Table 6. It is seen that larger self weight is
shared by the sheet-piles on the cases of either
the sheet-pile length become longer or the
ground became denser.
Predominant period of the ground-structure
system analyzed by the transfer functions (A/D
in Fig 7) during a small event (white noise
motion; Peak Acceleration=0.02g in prototype)
are summarized in Table 7. This shows the
initial stiffness of the SPF is larger than that of
the shallow foundations, and mainly influenced
by the ground condition.
Skelton curves, which consists of relations

CENTRIFUGE TESTS

1) Outline of the centrifuge models


Because of the stress dependent physical
properties of ground materials, centrifuge shake
table tests were conducted to simulate actual
characteristics of the SPF during earthquake
event. Centrifuge tests were carried out under a
25g centrifugal acceleration, therefore, the
model structures were scaled as 1/25, as
shown in Fig 7. A 9m high viaduct with 5m
square footing was selected as the prototype
(Photo 2). Sheet-pile was also scaled in this
experiment except connections as shown in Fig

Rigid Container

Strain Guages
SP40+pier2=42
Dry Sand

200 (5m)

250
11@16=176

Unit: mm

1900 (47.5m)

200
198
11

200
D

Fig 7(a): Centrifuge setup


(Two models were installed
in one model ground)

Shake Direction

100 (2.5m)

Earth Pres. Cell

300 200 (5m)

Disp. Transducer 2

11

Sheet-pile type III


t=13mm

400

Sheet-pile

t=0.5mm
10

15

50

Accelerometer

Shake Direction

400 (10m)

600 (15m)

in the limited area shallower than 50mm on


both sides of the footing. It is seen that the
deformation of the ground around the footing
spreads outward direction. On the other hand,
in the case of the SPF (Case-HL2), the
horizontal deformation of the ground is
restrained by the sheet-piles.

s=1/25
16

Fig 8: Sheet-pile model

Front/Back
Sheet-pile

Filled with
epoxy resin

Footing

Side Sheet-pile

Photo 2: Model structure and ground

Fig 7 (b): Plan view of the footing

249

ground. y of the SPF with L/B=1.0 is 0.6g in


this case, and this is twice as large as that of
the shallow foundation, of which the y is 0.3g,
as proportional to the dense ground cases.
Time histories of the earth pressure cells
installed at the bottom surface of the footings
during the severe earthquake input motion
(Bedrock PA=0.45g) event are shown in Fig 10.
From this figure, it is confirmed that relatively
large area under the footing is separated from
the ground surface during the earthquake event,
and the separation was recovered after the
earthquake.
Fig 11 shows the stress time histories of the
sheet-pile installed perpendicular to the shake
direction during the shake event mentioned
above. Both axial and bending stress
components are far smaller than its yield limit

between the horizontal response acceleration


(g) and the response displacement of the pier
top (mm), on the shallow foundation and the
sheet-pile foundations (L/B=1.0 and 0.5) at the
dense ground are shown in Fig 9. These
curves are found with sinusoidal input motion
tests. Yield acceleration y of the SPF with
L/B=1.0 is 0.8g, and this is twice as large as
that of the shallow foundation on the same
ground, of which y is 0.4g. y of the SPF with
L/B=0.5 is also improved as 0.6g. This means
with installing the sheet-piles in the ground,
seismic performance of the shallow foundation
can be largely improved, and the performance
is affected by the length of the sheet-piles.
Yield accelerations y are also compared in
Table 8, in the case of the medium dense

Table 4: Summary of test conditions


Case
Foundation type
Ground condition

1-1
Shallow
foundation

1-2

1-3

SPF

SPF

2-1
2-2
Shallow
SPF
foundation
Medium denseDr=60%,
Vs=180m/s
1.0B

Dense Dr=90% : Vs=200m/s

Sheet-pile length
Note : B is the footing width

1.0B

0.5B

Table 5: Summary of input motions for the shake table tests


Item
Sinusoidal motion
Earthuake motion

Summary
3 different frequencies (f=1.2, 2.4 and
3.6 Hz with 25 cycles)
A standard seismic motion, L2-spectral
1, prepared for the raylway structures
by RTRI

Purpose
Investigating the basic dynamic
characteristics of the SPF
Investigating the performance of the
SPF during the earthquake event

Table 6: Vertical self load distribution after the centrifuge acceleration process
Case
Foundation type
Ground
Sheet-pile
Vertical pressure
(kPa)
Vertical load
distribution
(Footing : SP)

1-1
SF
-

1-2
SPF
Dense
1.0B

1-3
SPF
0.5B

2-1
2-2
SF
SPF
Medium dense
1.0B
-

300

160

200

300

130

1 : 0.87

1 : 0.5

1 : 1.3

Note : SF is the shallow foundation; SP is the sheet-pile

Table 7: Predominant period of ground-structure system measured at the small shake event
Case

1-1

Foundation type
Ground
Sheet-pile
Predominant Period (s)

SF
0.66

1-2
SPF
Dense
1.0B
0.50

250

1-3

2-1

SPF

SF
SPF
Medium dense
1.0B
0.68
0.55

0.5B
0.51

2-2

MSP: Resistant moment against the rotation of


the footing by sheet-piles
Ms: In-plane bending moment by the side sheetpiles
NspC: Axial force of the sheet-pile at the front
side of the footing (Compression side)
NspT: Axial force of the sheet-pile at the back
side of the footing (Tension side)
MoutC: Bending moment of the front side sheetpile
MoutT: Bending moment of the back side sheetpile
B: Width of the footing
On the other hand, the rotation resistance
from the ground reaction under the footing
against the rocking motion is estimated by
assuming all vertical components can be

(240MPa) during the severe earthquake event.


This means the sheet-pile is not a critical
member on the seismic design of the SPF.
3) Bearing mechanism of the SPF
As it was seen, seismic performance of the
SPF is largely improved compared with the
shallow foundation. Therefore, the bearing
mechanism of the SPF against the seismic
action will be discussed in this section.
Bearing mechanism against the rocking
motion of the SPF is investigated utilizing the
force equilibrium around the footing defined as
eq.2, as illustrated in Fig 12.

MSP = Ms + (Nsp Nsp ) B / 2 + (Mout + Mout )


T

(2)

Earth Pres. (kPa)

1
Yield Acc. L/B=1.0
0.8
0.6 Yield Acc. of L/B=0.5
0.4
0.2
0
Yield Acc. of
-0.2
shallow foundation
-0.4
Shallow foundation
-0.6
SPF L/B=1.0
-0.8
SPF L/B=0.5
-1
-400
-200
0
200
400
Disp. (mm)

300
100
0
-100
0

50

Axial stress
Bending stress

50
0
-50
-100
0

2-2
SPF
1.0B
0.6

10

20
30
Time (sec)

40

Fig 11: Stress time histories of sheet-pile


mg

Mg

MoutC

NspT

40

100

Ms
MoutT

20
30
Time (sec)

Stress monitoring location

Table 11: Yield accelerations


(Medium dense ground)
2-1
SF
0.3

10

Fig 10: Earth pressure time histories

Fig 9: Load-displacement skelton curves of


tested foundations (Dense ground)

Case
Foundation type
Sheet-pile length
y (g)

Center
South 1/4

200

Stress (MPa)

Response Acc. (g)

Here,

NiS

NspC

Ng
b

NiT

Fig 12: Force equilibrium around the footing in


terms of rotation from the sheet-piles

251

NiC

Fig 13: Schematic drawing of the rotation


resistance from the ground reaction

50

supported by the vertical resistance of the front


side sheet-pile as its axial force, as shown in
Fig 13. Therefore, total vertical force from the
ground reaction Ng and the resistant moment by
the ground Mg can be calculated as follows.
(3)
N g = mg N i

assumption was suitable.


Since the calculation assumption of the
bearing mechanism is reasonable, share of
each reaction component against rotation can
be analyzed. According to the calculation
example of Fig 14, about 1/3 of the resistant
moment is shared by Mg (Ground reaction
component), and 2/3 is shared by MSP (Sheetpile reaction component).
Detailed analysis on the sheet-pile reaction
component is carried out as shown in Fig 15. At
the time step shown in the Fig 15(a), it is found
that 86% of the resistant moment is shared by
the axial resistance of the front/back sheet-piles,
and only 5% is shared by the bending
resistance of the sheet-piles. Share of each
component is illustrated in Fig 15(b).

Here,
mg: Mass of the superstructure x Centrifugal
acceleration
N i : Sum of the vertical force of sheet-piles
(4)

Here,
b: B/2
Therefore, total resistant moment against the
rotation around the footing is defined as eq.5.
M RSPF = M SP + M g
(5)

FULL-SCALE FIELD TEST

a) Outline of full-scale models


Aiming at practical use, the full-scale field
tests were conducted in Kawagoe-City, Japan.
The test setup and the models are shown in
Photo 3. The surface diluvial clay (Kanto loam),
with a thickness of 5m, is laid on the gravel
layer. These models have a 3.6m square
footing, 6m high pier. The sheet-piles length is
3.6m, the same as the width of the footing.
Therefore, the tips of the sheet-piles were not
installed into the gravel layer.

Rotation moment (kNm)

Case 1-2:L/B=1.0
Sin 1.2Hz, 0.4g

M
2
1

M (N)
out
M (S)

0.5
0.0

out

T=4.026sec
4.01

4.02
4.03
Time (sec)

4.04

4.05

Fig 15(a): Time histories of each resistant


moment component of MSP

SP

M +M
SP

1.0

Case 1-2:L/B=1.0
Sin 1.2Hz, 0.4g

-0.5
4.00

4
3

1.5

MoutT+ MoutC

Rotation moment (kNm)

2.0

* B/2

Resistant moment is calculated by eq.2


utilizing the forces measured by strain gauges
on the sheet-piles. The resistant moment by the
ground Mg is calculated at limited time section,
as illustrated in Fig 14. Calculation results show
there is good matching between MRSPF and MA.
Therefore, it is confirmed that the calculation

MSP

Time histories of the action moment


calculated from the inertia force of the structure
and the resistant moment defined as eq.5 are
shown in Fig 14. Action moment around the
footing center MA is calculated as follows.
M A = miili
(6)
Here,
mi: mass of the superstructure, pier and footing
i: Measured response acceleration
li: Arm from the center of the footing

(NSPT-NSPC)

M g = Ng b

Axial resistance by the front/back sheet-piles

Bending resistance
resistance
66of the front/back sheet-piles

-1

Rotation resistance by the side sheet-piles

-2

-4
3.7

0.09 0.05

Resistant moment from the ground Mg

-3

0.86

is calculated throughout this section only.


3.8

3.9

4.0
4.1
Time (sec)

4.2

4.3
0.0

0.2

0.4

0.6

0.8

1.0

Fig 15(b): Share of each resistant moment


component of MSP

Fig 14: Time histories of the action moment and


the resistant moment

252

condition.
3 Sheet-pile is not a critical member on the
seismic design of the SPF.

b) Horizontal static loading test


The horizontal static loading tests were
conducted by pulling the tops of two models to
each other by a hydraulic jack. At first, the
shallow foundation was pulled by the sheet-pile
foundation. Next the sheet-pile foundation was
pulled by the shallow foundation reinforced by
the ground anchor.
The P- relation of each case is shown in
Fig 16, together with the photos of the last
deformation of each model. Fig 16 shows that
the ratio of the horizontal resistance of the
sheet-pile foundation against that of the shallow
foundation was about four. This ratio is larger
than that obtained by the laboratory test.

Based on the above test results, a guideline


of design and construction of the sheet-pile
foundation was published by the Railway
Technical Research Institute, Japan, and
several actual projects were realized so far.
Because of its excellent performance and cost
competitiveness, application of the sheet-pile
foundation may increase in future.
REFERENCES
Koda, M., et al. (2003) The Proposal of Sheet-Pile
Foundation Combining a Footing with SheetPiles (in Japanese), TSU-CHI-TO-KISO, Vol. 51,
No. 11, pp.8-10
Nishioka, H., et al. (2004) A Series of Static Loading
Tests of Modeled Sheet-Pile Foundation
Combining Footing with Sheet-Piles on Sand,
Proceeding
of
15th
Southeast
Asian
Geotechnical Society Conference, pp.199-204,
Bangkok, Thailand.

CONCLUSION

Followings were found from a series of tests.


1 The seismic performances of the sheet-pile
foundation are greatly improved compared with
those of the shallow foundation.
2 Performance of the sheet-pile foundation is
influenced by both sheet-pile length and ground
Hydraulic Jack

Shallow
foundation

Sheet-Pile
foundation

Photo 3: Test setup and full-scale models

Horizontal load P (kN)

1000
800
600
Shallow Foundation
400

Sheet-Pile Foundation

200
0
0

100

200

300

400

500

Horizontal displacement (mm)

Fig 16: P- relationship and pictures of the models after loading

253

Rocking Seismic Isolation of Bridges Supported by Spread Foundations


K. Kawashima1, T. Nagai2, D. Sakellaraki2
1

Tokyo Institute of Technology, Japan


Former Graduate Student, Tokyo Institute of Technology, Japan

Abstract
This paper presents an analysis on the effectiveness of rocking seismic isolation of bridges
supported by spread foundations. Separations of footing from and contacts of footing with the
underlying ground which occur during an extreme ground motion result in mitigation of bridge
response. Separations and contacts of footing must have occurred in past earthquakes although
their effect was not rigorously included in seismic design. The effect of rocking seismic isolation is
presented for a 10 m tall standard bridge supported by spread foundations under three directional
excitation. It is shown that the effect of rocking seismic isolation is significant in reducing the plastic
deformation of columns at the plastic hinge regions although this increases deck and columns
response displacement.

INTRODUCTION

Direct or spread foundations are widely used


to support bridges where soil condition is stable.
They are designed so that performance for
sliding, settlement and overturning is assured.
Since overturning is generally critical in spread
foundations, it is important to clarify the safety
for overturning. Rocking response of spread
foundations provides a unique structural
response of bridges. Although spread
foundations have been conservatively designed
so that uplift from the underlying ground can be
minimum for preventing overturning, it is often
observed
from
the
post
earthquake
investigation that cracks along footings
occurred on the ground surface. This obviously
shows that rocking response occurred in spread
foundations during past earthquakes. Positive
use of rocking response of spread foundations
results in the isolation effect on bridges
response. The rocking response of spread
foundations increases as size of footings
decreases however it could result in overturning
of bridges under an extreme ground motion.
Consequently careful evaluation on size of
spread foundation is required.

254

Rocking response of structures has been


investigated by many researchers (Housner
1963, Ishiyama 1982, Kawashima et al. 1989,
Priestley et al. 1996). Housner studied rocking
response of a rigid block on a rigid base
(Housner 1963). Inelastic rocking response of
large rigid foundations was investigated by
Kawashima et al., and it was found that rigid
foundations did not overturn even if they were
subjected to ground motions with peak
accelerations
much
larger
than
static
accelerations in the static analysis (Kawashima
et al. 1989, 1991, 1994). This fact was taken
into account in the seismic design of HonshuShikoku Bridges, including the world longest
Akashi Straight Bridge and Kurushima Straight
Bridge.
More recently, Ciampoli et al. presented an
importance of rocking response and bridge
response interaction (Ciampoli et al., 1995).
Priestley et al. presented a contribution of
rocking response of a footing to a total deck
displacement
(Priestley
et
al.
1996).
Kawashima and Hosoiri (2003) and Kawashima
et al (2005) showed seismic rocking isolation
effect on standard bridges. Mergos and
Kawashima (2005) showed three directional
excitation effect on seismic rocking isolation of
a bridge (Mergos and Kawashima, 2005).

bottom of footing; FL and FLa : demand and


capacity of sliding force, respectively; e and ea :
eccentricity resulting from lateral force and its
allowable value, respectively; M B and V :
moment resulted from lateral force and vertical
force resulted from dead weight of deck,
column,
footing
and
overburden
soil,
respectively; and A : footing base area. It is
noted that ea is generally one third of
foundation width.
However, as rocking response increases,
footing starts to uplift and separate from the
ground at an edge as shown in Fig 1. Static
settlement of a spread foundation, v FS , with a
width l resulted from a static dead load of deck,
column and footing, V , may be written as

Sakellaraki and Kawashima verified seismic


rocking isolation effect based on shake table
test (Sakellaraki and Kawashima, 2006).
This paper shows the rocking isolation effect
of a bridge supported by spread foundations,
and clarifies the effect of bilateral excitation on
seismic rocking isolation.
ROCKING OF SPREAD FOUNDATION

In static seismic design, spread foundations


are sized so that the seismic performance can
be assured for based on bearing capacity of the
ground, sliding and overturning of foundation as

SB =

a A

< S Ba ; S S =

e=

MB
< ea
V

FL
< S Sa ;
FLa

v FS =

V
k sv l

(2)

When foundation uplifts with an angle F


from the resting position, an edge of
foundation starts to separate from the
underlying ground as upward displacement of
footing at edge, l FS / 2 , becomes larger
than initial settlement v FS~. Rotational spring
stiffness of foundation, K F , may then be
written as

(1)

in which, S B and S S , S Ba and S Sa : safety


factors for bearing capacity and sliding,
respectively, and their allowable safety factors;
a : vertical stress capacity of the ground at the

l F
l/2
2
< v FS
l / 2 k sv ( x) x dx
~
2
K F =
l FS
Xl / 2 k sv ( x) x 2 dx
v FS
2

(a)

(3)

in which X represents distance from the


center of footing to a point where footing uplifts
to the original level of the ground.
To
represent
rocking
soil-foundation
interaction, an analytical model as shown in Fig
2 (a) is widely used for a bridge supported by
spread foundations. Rotational spring stiffness,
K F , idealizes the rocking soil-foundation
interaction. Assuming that subgrade reaction of
soils per unit area, k sv , is linear at the entire
response displacement range of footing,
rotational stiffness of foundation K F may be
obtained as

(b)

K F = l /l 2/ 2 k sv ( x) x 2 dx

(4)

in which x is distance from the center of


footing.
Because the idealization by Fig 2 (a) with a
linear rotational stiffness K F only represents
linear response of a spread foundation, an

(c)
Fig 1: Uplift of spread foundation from the
underlying ground; (a) Static equilibrium, (b)
Uplift at left edge and (c) Uplifted equilibrium

255

separation as shown in Fig 3. Consequently,


the restoring force of the i-th spring is written as
k (v Fi v FSi )
fVi = SVi
0

v Fi < v FSi
(5)
v Fi v FSi

in which k SVi : stiffness of the i-th soil spring,


vFSi : initial settlement of footing due to dead
weight of bridge, and v Fi : relative
displacement at the i-th soil spring between the
footing and the ground which is defined as

v Fi = v Fi vGi

(a)

(6)

where v Fi and vGi are vertical response


displacement of footing and the ground,
respectively, at the i-th soil spring.
Although it is not shown in this paper, the
underlying ground may yield due to rocking of a
spread foundation when the underlying soil
does not have sufficient strength. The
saturation of the restoring force of the
underlying ground can be incorporated in the
idealization by Fig 3 (Kawashima et al 1991,
Kawashima and Hosoiri 2003).
It is known that sliding and uplift of a rocking
body develops a complex interaction between
sliding, rocking and jumping (Ishiyama 1982).
Sliding is however restrained in the following
analysis because sliding is not critical in spread
foundations embedded in the ground. The
analytical model shown in Fig 2 (b) can be
easily extended to three-dimensional model.

(b)

Fig 2: Idealization of spread foundation; (a)


without separation, and (b) with separation of the
footing from underlying ground

Fig 3: Idealization of restoring force of underlying


ground by soil spring

analytical model as shown in Fig 2 (b) is used


here to take account of uplift and separation of
a spread foundation from the underlying ground.
It is assumed in the model that the i-th soil
spring has nonlinear restring force with a
stiffness of k SVi in compression and 0 in

TARGET BRIDGE
A bridge which is analyzed here is presented
in Fig 4. It is a 200m long five span continuous
bridge supported by two abutments and four
reinforced concrete columns. Abutments and

Fig 4: Target bridge

256

design code (Japan Road Association 2002).


Since the bridge and the soil condition are
almost uniform along the bridge axis, a
structural system consisting of a column, a
foundation and a tributary deck is analyzed
here. The seismic response in the longitudinal,
transverse and vertical directions is analyzed.
The spread foundation determined from static
seismic
design
assuming
0.2g
lateral
acceleration is 2 m thick, 6.5 m long in the
longitudinal direction and 7 m wide in the
transverse direction. Overturning is the most
critical requirement in sizing foundations.
A foundation, a column and a part of deck are
idealized by a three-dimensional discrete
analytical model as shown in Fig 5. The footing
was idealized by a rigid grid consisting of beam
elements. Four sides of footing in the
longitudinal and transverse directions are
referred hereinafter A- and B-sides and C- and
D-sides, respectively (refer to Fig 5). The soilfoundation interaction was idealized by
nonlinear soil springs as shown in Fig 3. Soil
spring stiffness was determined from subgrade
reaction of soils based on the design practice
(Japan Road Association 2002). Column at the
plastic hinge was idealized by 3D fiber elements,

Fig 5:
Idealization of spread foundationunderlying ground-column-deck

Acceleration (m/s2)

columns rest on six spread foundations.


Superstructure is consisting of a plate girder
deck supported by five fixed steel bearings.
Spread foundations are embedded in sandy
soils with the bottom of foundations resting on
gravels with N-value of the standard penetration
test over 50. Since the fundamental natural
period of the ground is less than 0.2 s, this site
is designated as a stiff site (ground group I)
according to the Japanese highway bridge
9.8
0
-9.8

Acceleration (m/s )

(a)
9.8
0
-9.8

Acceleration (m/s )

(b)
9.8
0
-9.8
0

10

15
Time (s)

20

25

30

(c)
Fig 6: Ground motion for analysis; (a) NS, (b) EW and (c) UD components of JMA Kobe Observatory
record during the 1995 Kobe, Japan earthquake

257

Acceleration (m/s )

20
0
-20

Acceleration (m/s )

(a)
20
0
-20

Displacement (m)

(b)
0.3
0
-0.3

Displacement (m)

(c)
0.1
0
-0.1

10

15
Time (s)

20

25

30

(d)
Fig 7: Response accelerations and displacements at the deck under unilateral excitation when uplift of the
footing is not taken into account; (a) longitudinal acceleration, (b) vertical acceleration, (c) longitudinal
displacement, and (d) vertical displacement

Displacement (m)

0.15
A-side
B-side
0

-0.15

(a)
Restoring Force (kN)

1500
A-side
B-side
0

-1500
0

10

15
Time (s)

20

25

30

(b)
Fig 8: Response of the footing at A and B sides under unilateral excitation when uplift of the footing is not
taken into account; (a) vertical displacement of the footing, and (b) reaction restoring force of the soil spring

258

10000

500

500

5000

0
-500

-1000
-0.05
0
0.05
0.1
Relative vertical displacement (m)

Moment (kNm)

1000
Restoring force (kN)

Restoring force (kN)

1000

-5000

-500

-1000
-0.05
0
0.05
0.1
Relative vertical displacement (m)

(a)

-10000
-0.01

0
Curvature (1/m)

0.01

(b)

Fig 9: Restoring force of a soil spring under unilateral excitation


when uplift of the footing is not taken into account; (a) A-side,
and (b) B-side

and linear beam elements with the cracked


stiffness elsewhere. Flexure strength and
ductility capacity of the column were determined
based on an empirical stress and strain relation
of confined concrete (Hoshikuma et al 1997,
Sakai et al. 2000) and reinforcing bars
(Menegotto et al. 1973). Foundation and deck
were assumed rigid and their masses were
lumped at their gravity centers.
Because effect of overburden soil is
important in analysis, it was assumed that
weight of the overburden soil was included in
the evaluation of static settlement v FS in terms
of V in (1) and (2), but it was disregarded in
evaluation of the inertia force.
Three-dimensional bridge response under
bilateral and vertical excitation was computed.
As shown in Fig 6, NS, EW and vertical
components of ground accelerations measured
at Kobe Observatory of Japan Meteorological
Agency (JMA Kobe record) during the 1995
Kobe, Japan earthquake was imposed to the
bridge in the longitudinal, transverse and
vertical directions, respectively. Response
under unilateral and vertical excitation was also
computed for clarifying effect of unilateral
excitation. NS and UD components were
imposed to the bridge in the longitudinal and
vertical directions, respectively, in this
evaluation. Rayleigh damping was assumed to
represent energy dissipation (Clough and
Penzien 1993). Damping ratio was assumed
0.05 for the first and second modes.

Fig 10: Moment vs. curvature


hysteresis of column at the
Plastic hinge

For clarifying effect of three directional


excitation, response of the bridge under
unilateral and vertical excitation was first
evaluated. Fig 7 shows deck accelerations and
displacements when the bridge was subjected
to NS and US components in the longitudinal
and vertical directions, respectively. Soilfoundation interaction was idealized by a linear
rotational soil spring (refer to Fig 2(a)) without
taking uplift and separation of the footing from
the underlying ground into account in analysis.
The peak deck acceleration and displacement
in the longitudinal direction are 13.1 m/s2 and
0.219 m, respectively.
Fig 8 shows relative vertical displacement of
footing vFi and restoring force of a soil spring
fVi at both sides in the longitudinal direction (Aand B-sides, refer to Fig 5). The peak vertical
displacement of footing resulted from rocking
response was 15.2 mm and 22.9 mm at A-and
B-sides, respectively. Fig 9 shows restoring
force fVi vs. relative vertical displacement v Fi
hysteresis of the same soil springs. The peak
compression and tension of a soil spring is 903
kN and 472 kN, respectively, at A-side. They
result in compression and tension stress of 2.54
MPa and 1.33MPa, respectively, in the
underlying ground at A-side.
As a consequence, column shows moment
vs. curvature hysteresis at the plastic hinge in
the longitudinal direction as shown in Fig 10.
Column undergoes significant plastic range with
the peak curvature of 9.4x10-3/m.
On the other hand, Fig 11 shows deck
accelerations and displacements of the bridge

SEISMIC RESPONSE UNDER UNILATERAL


AND VERTICAL EXCITATION

259

Acceleration (m/s )

15
0
-15

Acceleration (m/s )

(a)
20
0
-20

Displacement (m)

(b)
0.3
0
-0.3

Displacement (m)

(c)
0.1
0
-0.1

(d)
Fig 11: Response accelerations and displacements at the top of column under unilateral excitation
when uplift of the footing is taken into account; (a) longitudinal acceleration, (b) vertical
acceleration, (c) longitudinal displacement, and (d) vertical displacement

Displacement (m)

0.15
A-side
B-side
0

-0.15

(a)
Restoring force (kN)

1500
A-side
B-side
0

-1500
0

10

15
Time (s)

20

25

30

(b)
Fig 12: Response of the footing at A and B sides under unilateral excitation when uplift of the footing
is taken into account; (a) vertical displacement of the footing, and (b) reaction restoring force of the
soil spring

260

750
0
-750

-1500
-0.05
0
0.05
0.1
Relative vertical displacement (m)

(a)

1500

4000

750

2000
Moment (kNm)

Restoring force (kN)

Restoring force (kN)

1500

0
-750

-2000

-1500
-0.05
0
0.05
0.1
Relative vertical displacement (m)

-4000
-0.001

0
Curvature (1/m)

0.001

(b)

Fig 13: Restoring force of a soil spring under unilateral excitation


when uplift of the footing is taken into account; (a) A-side, and
(b) B-side

Fig 14: Moment vs. curvature


hysteresis of column at the plastic
hinge

springs at A- and B-sides. The peak


compression of a soil spring is 725 kN at A-side,
which corresponds to compression stress of
2.04 MPa. No tension stress is induced in the
soil springs. The compression stress of 2.04
MPa is 80 % of the stress in the underlying
ground by disregarding uplift and separation of
footing from the underlying ground.
Fig 14 shows moment vs. curvature
hysteresis of column at the plastic hinge in the
longitudinal direction. The plastic deformation of
column is limited and the peak curvature is
7.63x10-4/m. This is only 8% the curvature
computed by disregarding uplift and separation
of footing from the underlying ground. Rocking
seismic isolation effect is thus significant for
mitigating damage of column at the plastic
hinge.

computed by idealizing nonlinear soilfoundation interaction by nonlinear soil springs


(refer to Fig 2 (b)) taking uplift and separation of
footing from the underlying ground into account.
The bridge was subjected to NS and US
components of JMA Kobe ground motion in the
longitudinal and vertical directions, respectively.
The peak deck acceleration and displacement
in the longitudinal direction are 6.69 m/s2 and
0.253 m, respectively, which are 49% smaller
and 16% larger, respectively, than those
computed by disregarding uplift and separation
of footing from the underlying grounds. It is
obvious that the significant decrease of the
peak deck acceleration results from rocking
seismic isolation. It should be assured that the
slight increase of deck displacement does not
result in any problem for the seismic
performance of bridge.
Fig 12 shows vertical relative displacement
of footing vFi and restoring force of a soil
spring fVi at A- and B-sides in the longitudinal
direction. Footing rocked and uplifted 110 mm
and 56 mm at A- and B-sides, respectively,
from an initial settlement due to dead weight
vFS (refer to (2)) of 2.48 mm. It is important to
note that spread foundation which was
designed in accordance with static design
assuming 0.2 g lateral acceleration by (1) uplift
56-110 mm under a near-field ground motion
recorded during the 1995 Kobe earthquake.
This means that rocking isolation must have
occurred in past significant earthquakes
although this effect was not considered in
design.
Fig 13 shows restoring force fVi vs. relative
vertical displacement vFi hysteresis of soil

SEISMIC RESPONSE UNDER THREE


DIRECTIONAL EXCITATION
Fig 15 shows deck responses under three
directional excitation evaluated by taking uplift
and separation of footing from the underlying
ground into account. The peak deck
acceleration and displacement are 7.16 m/s2
and 0.263, respectively, in the longitudinal
direction and 7.50 m/s2 and 0.194 m,
respectively, in the transverse direction.
Because the peak deck acceleration and
displacement in the longitudinal direction are
6.69 m/s2 and 0.253 m, respectively, under
unilateral and vertical excitation, the peak deck
acceleration and displacement increase by 7%
and 4%, respectively, under three directional
excitation. Fig 16 shows relative vertical
displacement of footing v Fi and restoring

261

Displacement (m)

0.3
0
-0.3

Displacement (m)

(a)
0.3
0
-0.3

Displacement (m)

(b)
0.1
0
-0.1
0

10

15
Time (s)

20

25

30

Displacement (m)

(c)
Fig 15: Response displacements at deck under three directional excitation when uplift of the footing is taken
into account; (a) longitudinal displacement, (b) transverse displacement, and (c) vertical displacement
0.15
A-side
B-side

-0.15

Displacement (m)

(a)
0.15

C-side
D-side

-0.15

Restoring force (kN)

(b)
1000
A-side
B-side

0
-1000

Restoring force (kN)

(c)
1000
C-side
D-side

0
-1000 0

10

15
Time (s)

20

25

30

(d)
Fig 16: Response of footing under three directional excitation when uplift of footing is taken into account; (a)
vertical footing displacement at A and B-sides, (b) vertical footing displacement at C and D-sides, (c) soil
spring reaction restoring force at A and B-sides, and (d) soil spring restoring force at C and D-sides

262

1000
Restoring force (kN)

Restoring force (kN)

1000
500
0
-500
-1000
-1500
-0.1

0.1

500
0
-500
-1000
-1500
-0.1
0
0.1
0.2
Relative vertical displacement (m)

0.2

Relative vertical displacement (m)

(a)

(b)

5000

5000

2500

2500

Moment (kNm)

Moment (kNm)

Fig 17: Restoring force of a soil spring under three directional excitation when
uplift of the footing is taken into account; (a) AC-corner, and (b) BC-corner

0
-2500
-5000
-0.001

0
Curvature (1/m)

0
-2500
-5000
-0.001

0.001

(a) Longitudinal direction

0
Curvature (1/m)

0.001

(b) Transverse direction

Fig 18: Moment vs. curvature hysteresis of column at the plastic hinge
under three directional excitation when uplift of the footing is taken into account

force of a soil spring fVi at A- and B-sides in


the longitudinal direction and C- and D-sides in
the transverse direction. The peak relative uplift
v Fi is 127 mm and 79 mm at A- and B-sides,
respectively. Because v Fi at A- and B-sides
under unilateral and vertical excitation is 110
mm and 56 mm, respectively, v Fi under three
directional excitation is 15% and 41% larger.
Fig. 17 shows the restoring force fVi vs.
relative vertical displacement v Fi hysteresis of
the soil springs at the corner of A- and C-sides
(AC corner) and the corner of B- and C-sides
(BC corner). The peak soil compression is 647
KN and 1107kN at the AC and BC corners,
respectively, which results in the peak soil
stress of 1.8 MPa and 6.23 MPa, respectively.
Consequently, the compression stress induced
in the underlying soil at the corners under three
directional excitation is nearly 1.8-2.2 times the
soil stress under unilateral and vertical
excitation. Protection of soils at the corners is

needed depending on bearing capacity of the


ground.
Fig 18 shows moment vs. curvature
hysteresis of column at the plastic hinge in the
longitudinal and transverse directions. The peak
curvature is 8.24x10-4/m and 2.0x10-4/m in the
longitudinal
and
transverse
directions,
respectively. Since the strength of column is
much higher in the transverse direction than the
longitudinal direction, the curvature of column in
the transverse direction is limited. Because the
peak curvature in the longitudinal direction is
7.63x10-4/m under unilateral and vertical
excitation, it increases by 7% under three
directional excitation.
CONCLUSIONS
Effectiveness of the rocking seismic isolation
of spread foundations on the seismic

263

Clough R.W. and Penzien J. (1993) Dynamics of


Structures, 2nd edn, McGraw Hill, New York.
Hoshikuma, J. Kawashima, K. Nagaya, K. and A. W.
Taylor (1997) Stress-Strain Model for Confined
Reinforced Concrete in Bridge Piers, Journal of
Structural Engineering, ASCE, 123(5), pp. 624633.
Housner, G. W. (1963) The Behaviour of Inverted
Pendulum Structures during Earthquakes,
Bulletin Seismological Society of America, 53,
pp. 404-417.
Ishiyama, Y. (1982) Motion of Rigid Bodies and
Criteria for Overturning by Earthquake
Excitations, Earthquake Engineering and
Structural Dynamics, 10, pp. 635-650.
Japan
Road
Association
(2002)
Design
Specifications of Highway Bridges, Maruzen,
Tokyo.
Kawashima, K. and Unjoh, S. (1989) Rocking
Response of a Rigid Foundation subjected to
Seismic Excitation, Civil Engineering Journal,
32 (10), pp. 60-66 (in Japanese).
Kawashima, K. and Unjoh, S (1991) Overturning of
Rigid Foundation Resting on Ground with
Insufficient Yield Strength, Civil Engineering
Journal, 33(3), pp. 54-59 (in Japanese).
Kawashima, K., Unjoh, S. and Mukai, H. (1994)
Inelastic Rocking of Direct Foundation during
an Earthquake, Civil Engineering Journal, 36(7),
pp. 50-55 (in Japanese).
Kawashima, K. and Hosoiri, K. (2003) Rocking
Response of Bridge Columns on Direct
Foundations, Proc. fib-Symposium, Concrete
Structures in Seismic Region, Paper No. 118
(CD-ROM), Athens, Greece.
Kawashima, K., Watanabe, G., Sakeraraki, D. and
Nagai, T. (2005) Rocking Isolation of Bridge
Foundations, 9th World Seminar on Seismic
Isolation, Energy Dissipation and Active
Vibration Control of Structures, pp. 609-630,
Kobe, Japan
Mergos, P.E. and Kawashima, K. (2005) Rocking
Isolation of a Typical Bridge Pier on Spread
Foundation, Journal of Earthquake Engineering,
9(2), pp. 395-414.
Menegotto, M.and Pinto, P.E. (1973) Method of
Analysis for Cyclically Loaded R.C. Plane
Frames including Changes in Geometry and
Non-Elastic Behavior of Elements under
Combined Normal Force and Bending, Proc.
IABSE Symposium on Resistance and Ultimate
Deformability of Structures Acted on by Well
Defined Repeated Loads, pp. 15-22.
Priestley, N. M. J., Seible, F. and Calvi, G. M. (1996)
Seismic Design and Retrofit of Bridges, John
Wiley & Sons, New York.
Sakai, J. and Kawashima, K. (2006) Unloading and
Reloading Stress-Strain Model for Confined
Concrete, Journal of Structural Engineering,
ASCE, 132(1), pp. 112-122.

performance of bridges was studied based on


nonlinear analysis for a 10 m tall standard
bridge. Analytical model which takes account of
uplift and separation of footing from the
underlying ground was verified by a shake table
test. Based on the results presented herein, the
following conclusions may be deduced on the
effect of rocking seismic isolation:
1) If separation of footing from the underlying
ground due to rocking response occurs, the
plastic deformation of column at the plastic
hinge significantly decreases as a result of
softening of moment vs. rotation hysteresis of
footing. As a consequence, inelastic rocking of
footing results in an isolation effect on the
response of bridge. However because the
isolation effect results in an increase of
response displacement of bridge, this must be
properly considered in design.
2) Spread foundation designed in accordance
with static seismic analysis assuming 0.2g
lateral static acceleration and the working stress
design approach may rock separating from the
underlying ground under near-field ground
motions. Uplift at the edge of footing is 80-130
mm in a 10 m tall standard bridge under JMA
Kobe ground motion in the 1995 Kobe, Japan
earthquake. This means that rocking seismic
isolation must have occurred in past significant
earthquakes although this effect was not
considered in design. Positive use of the
rocking seismic isolation brings benefit in the
seismic design of bridges.
3) Uplift and separation of footing from the
underlying ground can be analyzed using
nonlinear soil spring idealization as shown in
Figs 2 (b) and 3.
4) Three directional excitation results in an
increase of bridge response acceleration and
displacement. In particular, stress induced in
the underlying ground at the corners
significantly increases under three directional
excitation. Underlying ground at the corners
needs to be protected, if necessary, for use of
rocking seismic isolation.
REFERENCES
Ciampoli, M. and Pinto, P. E. (1995) Effects of SoilStructure Interaction on Inelastic Seismic
Response of Bridge Piers, Journal of Structural
Engineering, ASCE, 121 (5): 806-814.

264

Sakellaraki, D. and Kawashima, K. (2006)


Effectiveness of Seismic Rocking Isolation of
Bridges based on Shake Table Test, 1st

European
Conference
on
Earthquake
Engineering and Seismology, Paper No. 364, pp.
1-10, Geneva, Switzerland

265

266

Deep Foundations

Effects of Limit State Performance of Steel Bearings on a Bridge Upper


Structure
K. Ohtomo1, Y. Sato1, M. Sakai1
1

Central Research Institute of Electric Power Industry, Japan

Abstract
The present study discusses the effect of bearing support performance on a bridge
superstructure during a strong ground motion. A hybrid seismic response experiment for a
steel bearing is conducted for this purpose. Slip-type hysteresis is identified for bridge axis
and transverse direction excitations arising from contact between pin and other members in
the shoe. The degree of superstructure response is also affected by such hysteresis,
particularly in bridge axis direction.

INTRODUCTION

42000
M

42000
M

16435

MMovable Shoe

16665

The effects of seismic limit state


performance of bearings on a whole bridge
system during strong ground motions have
been greatly concerned from past earthquake
damages. Although rubber bearings or seismic
isolation devices have replaced steel bearings
since the 1995 Hyogo-ken Nanbu earthquake,
they are still in use in existing bridges. This
indicates that the effects of steel bearing failure
on earthquake resistance of a bridge are a
crucial issue.
There are several physical models (Otsuka
et al, 1997; Yabe et al, 1997; Iemura et al,
1998) for steel bearing required in developing
dynamic seismic response analysis model of a
bridge system. However, a systematic nonlinear
model for steel bearing performance is yet to be
established and geometric behavior modeling
should be also discussed as well as material
strength in developing a nonlinear model.
The objective of the present study is then to
experimentally discuss the effect of near limit
state in regard to fixed steel bearing on seismic
performance of a total bridge system. For this
purpose, a hybrid seismic experiment is
performed to examine earthquake response
nature of a total bridge system associated with
nonlinear behavior of steel bearings.

75000

Fixed Shoe

unitmm

Fig 1: Target bridge


Table 1: Properties of superstructure of the target
bridge
Mass per Unit Lengtht/m
2

EIkNm

19.0
7

Vertical

1.9510

Transverse

1.9510

GJkNm

6
5

7.73910

TARGET BRIDGE AND ITS MODEL

Fig. 1 shows a target bridge consisting of


160m long three-span continuous decks
supported by four rectangular reinforced
concrete piers with one fixed bearing and three
movable bearing conditions. The superstructure
is assumed to have 3,000t in mass. As far as a

267

Table 2: Properties of superstructure of reinforced concrete column


1st Point

2nd Point

M1(kNm) 1(1/m) M2(kNm)


Fixed Shoe Column
Movable Shoe Column

-5

-5

Bridge Axis 1.5210 3.9010


Transverse 1.1510 3.9010

3rd Ponit

2(1/m)

1.3410 6.5510
1.0110 6.5510

M3(kNm) 3(1/m)
5

-3

-3

1.7210 3.8110
1.3010 3.8110

Bridge Axis,
4
-5
4
4
5
-3
1.0010 3.7010
Transverse 1.2710 4.9010 7.0610 7.5610

Table 3: Restraining of relative displacement between superstructure and substructure


Shoe Condition

Bridge Axis

Transverse

Vertical

Around
Bridge Axis

Around
Transverse

Around
Vertical

Fixed Support

Restrained

Restrained

Restrained

Restrained

Free

Free

Movable Support

Free

Restrained

Restrained

Restrained

Free

Free

HYBRID SEISMIC RESPONSE EXPERIMENT

bearing seismic performance is concerned,


major bridge damage due to poor bearing
performance
during
an
earthquake
is
predominant on three continuous girder bridges
among
other
bridges
(JSCE,
1996).
Furthermore,
such
bridge
structure
is
susceptible to sustain localized damage on the
piers that support fixed steel bearings. Then,
the three continuous girder bridge well satisfies
the present study purpose.
Three dimensional finite element seismic
response analysis (hereafter, 3DFEM analysis)
models of the target bridge were constructed
respectively for bridge axis and transverse to
bridge axis directions to facilitate numerical
analysis and experimental works in the present
study The superstructure properties (mass per
unit length, flexural rigidity EI and torsional
rigidity GJ) are tabulated in Table 1. Reinforced
concrete columns corresponding to both fixed
and movable bearings were idealized as a
beam element with a tri-linear moment M and
curvature relationship and their parameters
are listed in Table 2. Mass density of the
columns was set to be 2.4 t/m3.
The fixed and movable bearings were
modeled as just nodal in the 3DFEM analysis
and their freedoms are summarized in Table 3.
Although soil springs are often used to express
dynamic soil and foundation interaction in a
dynamic analysis, the current study discards
this option due to the limitation on degrees of
freedoms that can be handled in the hybrid
experiment control.

In fact, a hybrid experiment scheme


employed in the present study is a real-time
hybrid experiment that has an actuator
response delay compensation function using a
digital signal processor (DSP) instead of a
classic pseudo dynamic experiment procedure.
Hereafter, the term real-time is dropped unless
confusion arises.
Bearing
The present study concerns the earthquake
damage only on a bearing body. The pin-type
bearing specimen prepared in the hybrid
experiment is schematically illustrated in Fig. 2.
The pin is a steel cylinder whose middle part
diameter
is
intentionally
narrowed
for
assembling purpose with upper and lower
shoes. The metal material was SS400, which is
recommended in the Handbook of Bearing
Support for Highway Bridge (Japan Road
Association, 1991).
To design the pin profile, demand and
capacity balance on the transverse direction
excitation associated with the applied force F
was considered (Japan Road Association,
1991; Abe et al, 2004) as illustrated in Fig. 3.
The outer diameter D and the narrowed part
diameter d of the steel cylinder can be
determined by the tensile strength capacity of
the narrowed part and the shear strength
capacity at the shear key characterized by
thickness t in the shoe (Abe et al, 2001). Finally,
40mm and 20mm for the outer diameter and the

268

210

Pin
400

400

Pin

14.25

Transverse
Lower Shoe
Fig 2: Configuration and dimension of bearing specimen

Bridge axis

Acceleration(m/s )

8
4
0
-4
-8
0

112

Unitmm

8.25

14

Upper Shoe

10

105 105

4
25
25

400
20
40

Upper Shoe

400

10
Time(s)

15

20

Fig 4: Takatori ground motion (EW) component

Fig 3: Shear key structure of the pin

25
h0.05
2

Acceleration(m/s )

narrowed part diameters respectively were


determined so that the cylinder would sustain
breakout under the maximum load capacity of
the hydraulic actuator used in the hybrid
experiment, i.e., 500kN.
Numerical Model Development
An EW component of the observed ground
motion at Takatori-eki station of JR West during
the 1995 Hyogo-ken Nanbu earthquake
(Hereafter, Takatori record for simplicity) was
used for the dynamic analysis throughout this
study. In reality, first twenty seconds of duration
of Takatori record is used for analysis purpose.
The acceleration time history and the
acceleration response spectrum (h=5%; h
represents damping ratio) are presented in Fig.
4 and Fig. 5, respectively.
A nonlinear dynamic analysis was
implemented to assess earthquake response of
the target bridge. The equation of motion was
solved by numerical direct time integration
scheme, or Newmarks method ( =0.25) was
employed with time increment of 0.0007812s.
Rayleigh type damping was used to express
viscous damping constants involved in the
equation of motion. Modal damping ratio of 2%

20
15
10
5
0
0

2
Period(s)

Fig 5: Acceleration response spectrum

is assumed to be valid on the first and the third


natural periods for both bridge axis and
transverse directions.
The 3DFEM analysis model as constructed
previously for the target bridge was then
reduced to a two-degrees-of-freedom model
(hereafter, 2DOF model) for bridge axis as well
as transverse to bridge axis directions as
presented in Fig. 6. In order to develop the
2DOF model, the analytical correlation between
the 2DOF and the 3DFEM model analyses was
examined
focusing
on
the
maximum
displacement response at the column top under
Takatori record excitation.

269

Superstructure

Superstructure

Mass
Steel Bearing
Vertical Actuator (1000kN)

Stiffness
(Experiment)
Steel Bearing

Pier Mass
Horizontal Actuator (500kN)

Pier

Reaction Wall

Pier
Stiffness

Steel Bearing
Support Beam for Bearing

Strong Floor

Fig 6: 2DOF model for the hybrid experiment

Fig 7: Experiment setup (bridge axis direction)


Table 4: Dynamic properties of 2DOF model
Bridge Axis

Transverse

15.728

7.478

Superstructure
Mass(t)

To assess seismic response of the target


bridge, a similitude rule must be applied. The
similitude rule widely used among many hybrid
seismic response experiments for a bridge
structure is employed as presented in Table 5.
The scale factor was determined on the ground
that dimensions of the model bearing can be
scaled down from a so-called 300t class
bearing whose narrowed part diameter size is
equal to 60mm by referring one (Hanshin
Expressway Company Limited, 1989) of the
practical design examples. As a result, the
scale factor was determined to be 3.2. In fact,
five bearings are assumed to exist on the pier.

1.187

Pier Mass(t)
Pier Stiffness(kN/m)

1.4610

9.29104

Damping Ratio

0.2086

0.2825

Natural FrequencyHz

1.140

1.269

Natural Periods

0.8773

0.7879

Table 5: Similitude rule for the hybrid experiment


Quantity

Dimension

Scale Factor

Length

Mass

S3

Time

Velocity
Acceleration

-1

-2

1/S

LT

LT

Force

MLT-2

S2

Rigidity

MT-2

Experiment Setup
The schematic view of the hybrid
experiment setup is depicted in Fig. 7.
Photographs of transverse and bridge axis
directions setups are also shown Fig. 8 and Fig.
9, respectively. The horizontal and vertical
loads were applied to the bearing specimen by
500kN and 1,000kN actuators, respectively.
The 500kN actuator played a role on response
displacement loading from the computed
superstructure response, while the 1,000kN
actuator added the dead weight of the
superstructure.
The 500kN loading was applied to the
bearing specimen as a computed response
under Takatori record for bridge axis and
transverse directions. In other words, for an
example, the pin axis was placed perpendicular
to the 500kN actuator axis for representing a
bridge axis excitation.

The superstructure mass, the column mass


and the column stiffness were determined so
that the both displacement responses fall into
an acceptable tolerance. The dynamic
properties of the established 2DOF are
tabulated in Table 4. As far as the column
stiffness is discussed, values of equivalent
stiffness and damping ratio for the column
correspond to nonlinear load and displacement
relationship listed in Table 2 were identified
using an equivalent linear model approach.

270

Reaction Force (kN)

600
400

6.66m/s

8.00m/s

200
0
-200
-400
-600
-8

-4

0
4
Displacement (mm)

Fig 10: Hysteresis behavior of bearing (bridge axis)

Fig 8: Entire view of transverse direction setup

Upper Shoe

Pin

Lower Shoe

Fig 11: Translation movement around the pin

friction force between the upper and lower


shoes. The positive and negative reaction
forces correspond to the slip stays almost at
constant values regardless of the magnitude of
PGA. These absolute values are, however,
unequal between positive and negative
movement. This may depend on intact
inclination of the bearing seating.
The cause of relatively large stiffness as
observed in Fig. 10 can be explained as
follows: When the shoe rotates around the pin
axis, a slight translation of the pin occurs due to
existing gap as illustrated in Fig. 11. After the
translation reaches the pin-support surface
between upper and lower shoes, like a
mechanical key appears, resulting in essential
stiffness of bearing body.
Fig. 12 also presents load and displacement
hysteresis under 6.66m/s2 and 8.00m/s2 in PGA
excitations,
in
which
rotation-induced
displacement is extracted. Then, slip length is
obtained as about 0.5mm. In addition, the
difference between the pin diameter and the
pin-support diameter is 0.5mm, which indicates
that the maximum translation should be equal
to 0.5mm and meets the slip length as shown in
Fig. 12. This observation demonstrates the

Fig 9: Local view of axis direction setup


RESULTS AND DISCUSSIONS

Excitation Cases
Peak ground acceleration (PGA) value
equal to 6.66m/s2 in Takatori record was
adjusted like 1.11m/s2, 3.33m/s2, 6.66m/s2 and
8.00m/s2. Then, four sets of excitation cases
were applied for the bridge axis direction.
According to the similar manner with the case of
bridge axis direction, 1.11m/s2, 3.33m/s2,
6.66m/s2, 10.00m/s2 and 20.00m/s2 in PGA
were applied for the bridge transverse direction.
Hysteresis Characteristics of Bearing
Behavior
Fig. 10 shows the load and displacement
relationship under 6.66m/s2 and 8.00m/s2 in
PGA excitation cases for the bridge axis
direction. A slip-type hysteresis loop is clearly
observed in the respective case. The slip is
probably due to rotation of the bearing around
the pin axis, which simultaneously occurs when
the applied load exceed the maximum static

271

above-mentioned discussion.

Reaction Force (kN)

600
400

6.66m/s

8.00m/s

200
0
-200
-400

Fig 14: Breakout of the pin

-600
-2

-1

0
1
Displacement (mm)

Fig. 14 presents the pin breakout occurred


during the 20.00m/s2 excitation. One can
observe that the failure surface deviates 45
degrees from the pin axis. However, this seems
to be still brought by a tensile force. The tensile
strength of the narrowed part of the pin is
calculated about 160kN, this value virtually
becomes about 200kN because of additional
static friction force of 40kN. On the other hand,
the breakout load is identified as 247kN as
shown in Fig. 13, leading a twenty percent
increase to the above-mentioned value.

Fig 12: Hysteresis behavior of bearing using


corrected displacement (bridge axis)
300
Reaction Force (kN)

Estimated breakout load


200
2

6.66m/s
2
20.00m/s

100
0
-100

Estimated breakout load


-200

Effects on the Superstructure


Experimental results discussed here are
enlarged to a real size and correlated with the
2DOF analysis results. In this respect, the
2DOF model excludes the relative displacement
of the bearing to investigate clearly the effect of
the bearing performance.
The superstructure displacement time
histories for the bridge axis direction are plotted
in Fig. 15. The duration patterns are similar to
each other, especially peak response are
almost identical in the duration of 10s and more.
On the other hand, the magnitude of the peak
response in the experiment result is relatively
larger than those of in the analysis result. Thus,
the effect of the nonlinear hysteresis
characteristics of the bearing on the
superstructure response is observed. The
maximum displacement obtained by the
experiment is 18.4mm or about 14% larger than
that of the 2DOF analysis result. This suggests
that essential response of a bridge
superstructure needs to be estimated taking the
bearing hysteresis into account.
The slip-type hysteresis as shown in Fig. 10
undoubtedly contributes to such a response. As
a result, relatively larger peak responses occur

-300
-8

-6

-4
-2
Displacement (mm)

Fig 13: Hysteresis behavior of bearing (transverse)

Fig. 13 indicates load and displacement


relationship under 6.66m/s2 and 20.00m/s2 in
PGA excitations for the bridge transverse
direction, in which significant slip-type
hysteresis loops are also obtained. Note that
pin breakout at the narrowed part occurred
during the 20.00m/s2 excitation; subsequently
the hybrid experiment was terminated due to
the limiting values of displacement and force of
the 500kN actuator. The degree of the reaction
force at which the slip occurs is roughly
constant under respective excitation, indicating
that the pin moves along with the gap around
the shear key (See Fig. 3) of the bearing when
the applied force becomes larger than the static
friction force. The magnitude of the slip
movement is further intensified arising from the
enlargement of the narrowed part brought by
cyclic stress-induced plastic deformation.

272

reinforced concrete piers and a superstructure


are numerically modeled and assigned to the
numerical part and a steel bearing is dealt with
the experimental part. From the hybrid
experiment and additional numerical analysis
results, the following conclusions can be
addressed:
1) Load and displacement characteristics is
identified as a slip-type hysteresis for both
bridge axis and transverse to bridge axis
directions. In bridge axis direction, rotation of
the upper and lower shoes around the pin
contributed to develop such a specific loop. On
the other hand, the pin movement in the gap
distance around the shear key plays a role on
the slip-type hysteresis characteristic for
transverse direction. The magnitude of the slip
is further intensified under accumulated cyclic
loading.
2) The slip-type hysteresis indeed affected
the superstructure response. The maximum
horizontal displacement at the superstructure
for bridge axis direction and transverse direction
to bridge axis in the hybrid experiment are
larger than those of the 2DOF analysis that
uses a prescribed condition at the bearing to
represent the bearing behavior. Namely, 14%
and 4% increases for bridge axis direction and
transverse direction to bridge axis, respectively
are recognized. In particular, the 14% increase
may have a serious consequence on the design
of the gap allowance between adjacent decks.

in negative value in the reflection of negatively


biased hysteresis loop.
0.2
Displacement(m)

Experiment
2DOF

0.1
0
-0.1
-0.2

10
Time(s)

15

20

Fig 15: Superstructure displacement time histories


in bridge axis under 6.66m/s2 excitation
0.1
Displacement(m)

Experiment
2DOF

0.05
0
-0.05
-0.1

10
Time(s)

15

20

Fig 16: Superstructure displacement time histories


in bridge transverse under 6.66m/s2
excitation

ACKNOWLEDGEMENT

Fig. 16 compares the superstructure


displacement time histories for the bridge
transverse direction obtained by the hybrid
experiment and the 2DOF analysis. Peak
responses as well as duration characteristics
are
almost
identical.
The
maximum
displacement is 98.2mm and 94.2mm for the
experiment and the 2DOF analysis, respectively.
This comparison indicates that the experimental
maximum response is about 4% larger than that
of the 2DOF analysis and therefore the effect of
the
slip-induced
displacement
on
the
superstructure is insignificant.
COCLUDING REMARKS

The present study examined the effect of


the bearing performance associated with limit
state on a superstructure based on the hybrid
seismic response experiment in which

The present study was granted by NIED


(National Research Institute for Earth Science
and Disaster Prevention) in connection with FY
2005 US-Japan corroboration of experimental
studies on seismic performance of bridge
structures by utilizing Three-Dimension FullScale Earthquake Testing Facility (E-Defense)
program.
REFRENCES
Abe, M., Yanagino, K., Fujino, Y. and Hashimoto, S.
(2001) "Damage Analysis of Three-Span
Continuous Girder Bridges in 1995 Hyogo-ken
Nanbu Earthquake (in Japanese) ", Journal of
Structural
Mechanics
and
Earthquake
Engineering, No.668/I-54, pp.83-101
Abe, M., Yoshida, J., Fujino, Y., Morishige, Y, Uno S.

273

and Usami, S. (2004) "Experimental Investigation


of Ultimate Behavior of Metal Bridge Bearing
under Seismic Loading (in Japanese) ", Journal
of Structural Mechanics and Earthquake
Engineering, No.773/I-69, pp.63-78

JSCE Editorial Committee for the Report on the


Hanshin-Awaji Earthquake Disaster, (1996)
Report on the Hanshin-Awaji Earthquake
Disaster, Damage to Civil Engineering Structures,
Bridge Structures (in Japanese), pp.59-68

Hanshin Expressway Company Limited (1989)


Standard Design Drawings of Steel Bearing for
Steel Girder (in Japanese)

Otsuka, H., Kanda, M., Suzuki, N. and Kawakami, M.


(1997) "Dynamic Analysis Concerned with
Rotational Displacement of Skewed Bridges
Caused by Horizontal Ground Motion (in
Japanese) ", Journal of Structural Mechanics and
Earthquake Engineering, No.570/I-40, pp.315324

Iemura, H., Miyamoto, A. and Takahashi, Y. (1998)


"Influence of Failure of Steel Bearings on
Damage Modes of Bridges under Strong
Earthquake Motion (in Japanese) ", Journal of
Structural Engineering, Vol.44A, pp.659-666

Yabe, M., Takemura, H. and Kawashima, K. (1997)


"Effects of Impacts between Two Adjacent Decks
in a Straight Bridge and a Deck and Abutment in
a Skewed Bridge (in Japanese) ", Journal of
Structural Engineering, Vol.43A, pp.781-791

Japan Road Association, (1991) Handbook of


Bearing Support for Highway Bridge (in
Japanese)

274

Large-scale model tests of shallow foundations subjected to earthquake


loads
M. Shirato1, T. Kouno1, S. Nakatani1, R. Paolucci2
1

Public Works Research Institute, Japan


2
Politecnico di Milano, Italy

Abstract
We conducted a series of 1G large-scale shake table tests and cyclic eccentric loading tests
of a shallow foundation model. The experimental parameters were the difference in loading
methods (i.e. dynamic and static), input seismic motions (i.e. intensity and number of cycles), soil densities (i.e. dense and medium dense), and the ratio of horizontal and overturning moment loads. The result provided benchmark data sets for the development of numerical models for the response of a shallow foundation subjected to a large earthquake. The
experimental data set contained the acceleration and displacement of the soil and those of
the foundation model. The dataset also contained the distribution of reaction normal and
shear forces at the foundation base. The experimental results suggest that the coupling effect among vertical, horizontal, and overturning loads, the uplift of shallow foundation and a
sophisticated hardening rule should be considered to separately predict the evolution of the
residual displacement.

soil is caused by an increase in the eccentricity


of vertical loads especially when a shallow
foundation is rested on sturdy ground. It results
in a nonlinear relationship between a moment
and rotation of footing. The Specifications for
Highway Bridges has considered this kind of
force reduction in the structural design of shallow foundation footings for large earthquakes. A
nonlinear rocking load-displacement curve is
calculated using a Beam-on-Winkler Foundation
model. Then, the energy conservation rule can
be used to estimate the intensity and position of
the reaction force normal to the foundation base.
Recent case histories also back up the empirical engineering judgment that current shallow foundations have a sufficient safety margin
even against large earthquakes. The damage to
highway bridge shallow foundations such as
excessive settlement or inclination was not
caused in recent large earthquakes in Japan.
However, our understanding of the actual
behavior of shallow foundations during large
earthquakes is not sufficient enough to model
shallow foundations as macroscopic structural
elements in the seismic design calculation of

INTRODUCTION

Shallow foundations are considered to have


larger safety margins than pile foundations
against large earthquakes, because design
norms request that shallow foundations are directly rested on sturdy bearing layers. As for the
Specifications for Highway Bridges that is the
highway bridge design norm in Japan, further
empirical regulations for design of the normal
situation and small-to-mid scale earthquake
situation are required. For example, the choice
of bearing layer, the limitation of maximum soil
reaction stress intensity, and the limitation of
the degree of partial uplift are specified to prevent the foundations from an excessive settlement from the viewpoint of long-term serviceability. After all, such additional safety margins
can prevent bearing failure, excessive settlement, and inclination even during large earthquakes.
In addition, seismic loads to shallow foundations are considered to be reduced during large
earthquakes. It is expected that a progressive
reduction in contact area between footing and

275

structures. In design, we are unable to estimate


the plastic displacement caused by the cyclic
large force to the soil beneath the footing. Much
of the current design just owes the past experience.
The behavior of a shallow foundation subjected to a combination of vertical load (V),
horizontal load (H), and moment (M) has been
extensively investigated. As for the ultimate
bearing capacity of shallow foundations to
combined loading, a failure locus concept in the
V-H-M space has been developed (Nova and
Montrasio 1991, Butterfield and Gottardi 1994,
Houlsby and Martin 1993). As for the displacement of the foundation up to the ultimate state,
the macro-element approach has been developed (Nova and Montrasio 1991, Gottardi and
Butterfield 1995, and others). A work-hardening
plasticity is applied to the evolution of the yield
locus in the V-H-M space. As a combined load
increases, the yield locus expands and finally
becomes the failure locus in the V-H-M space.
Therefore, the foundation response can be obtained within the context of plasticity with a
relevant flow rule. This approach is much simpler than, for example, other sophisticated finite
element approaches and easy to implement
into the computation of dynamic soil-structure
interactions (Paolucci 1997, Cremer et al. 2001,
Okamura and Matsuo 2002, di Prisco et al.
2002).
However, the current macro-element modeling is basically based on monotonic loading experiments and the investigation of the macroscopic foundation behavior for cyclic combined
loading is nascent. Although, recently, some
experimental findings have shown the seismic
behavior of shallow foundations (e.g. Haya and
Nishimura 1998, Negro et al. 2000, Faccioli et
al. 2001, Gajan et al. 2005), available experimental datasets are still limited.
This paper presents the results of large
scale 1G tests of model pier footings on sand
that were subjected to shake table loading, cyclic lateral loading, and concentric vertical loading. This paper first introduces the testing
methods and then represents major findings
that should be useful for the establishment of
macro-element modeling for earthquake loading.
Note that a numerical simulation for the present
experiment is tackled in a companion paper
(Paolucci et al. 2007).

276

SHAKE TABLE EXPERIMENTAL PROGRAM

The experiments were conducted at the


Large-scale Shake Table Facility in the Public
Works Research Institute, Tsukuba, Japan. The
details in the experiment are reported in a
PWRI report (Fukui et al. 2007b).
Test apparatus

Fig. 1 shows photos of the experiment setup. The size of the shake table was 8 m 8 m
in plan. A laminar shear box having internal dimensions 4 m 4 m in plan and 2.2 m high was
placed on the table. It consisted of 10 layered
frames. The shake table was rocked only in the
North-South horizontal direction.
A dry dense sand deposit was made of Toyoura sand up to a height of 2.1 m in the laminar
shear box and compacted in layers so that a
satisfactory homogeneous soil condition was
obtained. The soil relative density, Dr, was 80%,
and the mass density, , was 1.60 103 kg/m3.
CD triaxial compression tests revealed that the
internal friction angle, , is 42.1. Undrained cyclic triaxial compression laboratory tests determined a small strain elastic modulus E0:
E0 = 34.92 'c0.4436 103 kN/m2

(1)

in which the unit of confining stress, 'c, is


kN/m2. 33 accelerometers were embedded in
the sand deposit.
A model pier footing was located on the center of the sand deposit surface. Fig. 2 shows a
schematic diagram of the model pier footing.
The model comprised of three main structural
components: a top steel rack, a short steel Ibeam column, and a footing. Fig. 2 also shows
the mass, mG, structural moment of inertia
about the gravity center, J, and height from the
footing base to the gravity center, hG, for each
structural component. The total weight of the
structural model was 8385 N. The total height of
the model was 0.753 m. The height of the center of gravity was 0.420 m from the base of the
footing. A steel rack at the top was 5227 N
heavy, including steel plates gathered on the
rack. A short steel beam having an I crosssection and end diaphragms connected the
steel rack and the footing. The I-beam connection was much stiffer than the soil-foundation
system and is considered rigid. The footing
shape was a 0.5 m sided square block. 11 bi-

earthquake are greatly affected by both intensity and duration (or the number of cycles) of
the earthquake, and, therefore, two types of
ground motion are involved in design. The Type
I seismic motion is ground motion associated
with the interplate-type earthquake having a
magnitude of approximately 8 and is generated
at plate boundaries in the ocean. The Type II
seismic motion is ground motion associated
with the inland-strike-type earthquake having
magnitudes of approximately 7 and is caused
by faults located at short distances from bridge
sites. The peak amplitudes of Type I seismic
motions are smaller than those of Type II seismic motions, but Type I seismic motions have
longer durations. Type II motions have high intensities but short durations. The shake table
was also excited by a sweep wave to check the
basic vibration properties at different stages.
The shake table motions were captured with
laser displacement transducers and accelerometers.

directional load cells were attached as the base


of the footing, so that the distribution of normal
and shear reactions to the base of the footing
was captured. The long side of the load cell had
the same length as the foundation side and it
was perpendicular to the excited direction. The
contact surfaces of the load cells to the soil deposit were rough, being covered with sand paper. 25 accelerometers were also attached to
capture horizontal and vertical accelerations of
the model pier footing.
The static safety factor in terms of the bearing capacity is considered to be 29, based on a
monotonic vertical loading experiment result
that will be mentioned later. It is much larger
than the required safety factors of the Japanese
highway design norm, a safety factor of 3 for
the normal design situation considering both
dead and live loads and 2 for the small-to-midscale earthquake design situation considering
dead loads and the effect of the earthquakeinduced inclination and eccentricity of load on
the bearing capacity. However, as mentioned in
INTRODUCTION, the Japanese design norm
also requires the limitations of the degree of
partial uplift and the intensity of subgrade reaction stress for both normal and small-to-midscale earthquake design situations. As a result
that we examined several past design results of
highway bridge shallow foundations, the ratios
of the bearing capacities to the dead loads
tended to be in the range of 9 to 24. The ratio of
the height of the center of gravity of the model
pier footing to the footing length was 0.84, and
it is also included in typical ratios in past design
case histories of highway bridge pier footings
(Fukui et al. 1999). Therefore, the model pier
footing had similar mechanical properties to
those of typical design cases.
Fig. 3 shows the two earthquake records
that rocked the shake table. A motion recorded
at Shichiho Bridge, Hokkaido, Japan, during the
1993 Hokkaido Nansei-Oki Earthquake (MW =
7.8) and the N-S component recorded at JMA
(Japan Metrological Agency) Kobe during the
1995 Hyogo-ken Nanbu (Kobe) Earthquake (MW
= 6.9) were adopted. The Shichiho Bridge motion is consistent with the Type I seismic motions and the Kobe motion is a very representative of the Type II seismic motions in the Japanese Specifications for Highway Bridges (JRA
2002). The characteristics of inelastic behavior
of both soils and structural members during an

Test series

We had two test series, Case S1 and Case


S2, where S denotes the shake table experiment. The sand deposit was separately prepared for each Case. Each case had several
excitation phases with different input motions.
Test series are tabulated in Table 1. Because
the system was gigantic, earthquake motions
were not perfectly reproduced on the shake table. Case S1 was originally planned for Case
S1-2, i.e. Type I seismic motion case, and Case
S2 was planned for Case S2-2, i.e. Type II
seismic motion case. However, after Case S1-3,
we rehearsed for the Case S2-2 experiment,
applying the Kobe motion several times with
different experimental conditions as test runs.
Although we were originally thinking that the
data was going to be dumped, the test excitation phases exhibited unique behavior, and we
ended up dealing with them as formal excitation
cases S1-4 and S1-5 and analyzing the data.
In Case S2-2, the Kobe motion was applied
but the acceleration amplitude was 80% of the
original record in time domain to prevent from
overturning failure and obtain a dynamic shallow foundation motion until the end of the excitation. Although a 10 mm embedment (4% of
the footing height) was involved, we presumed
we can disregard the embedment effect on the
overall foundation behavior.

277

Test apparatus
Data processing

Fig. 5 shows a schematic diagram and a


photo of the experimental setup. In a deep test
pit having internal dimensions 4 m 4 m in plan,
a dry sand deposit was made of Toyoura sand
to a height of 2 m with an average relative soil
density of Dr = 80% (soil density = 1.60 103
kg/m3) or 60% ( = 1.60 103 kg/m3). The deposit was compacted in layers so that homogeneous soil conditions were achieved. At a relative density Dr of 60%, CD triaxial compression
tests revealed that the internal friction angle, ,
was 39.9 and undrained cyclic triaxial compression laboratory tests determined a small
strain elastic modulus E0:

As shown in Fig. 4, the measured forces


and displacements of the footing are expressed
in terms of the displacements and resultant
forces at the base center of the footing, where v
is the settlement, u is the sliding, is the rotation, V is the vertical force, H is the horizontal
force, and M is the moment. The positive direction of the sliding, u, in Fig. 4 coincides with the
North direction in Fig. 2. The horizontal displacement of the model pier footing relative to
the soil deposit surface displacement was taken
and will be shown hereafter. Because the settlement of the soil deposit was little, the settlement will be shown using the absolute displacement hereafter.
In the following results, accelerations were
obtained with accelerometers. Displacements of
the model pier footing and soil deposit surface
were captured via an image analysis of digital
VCR records. The VCR image analysis was
able to capture more accurately the long-period
part of displacement, including permanent settlement and rotation. The double-integration of
acceleration together with a proper high-pass
filtering process also can provide displacement
records. Comparisons of the displacement records obtained with the VCR image analysis
and the double-integration of acceleration
showed that the phase changes and amplitudes
agree well with each other in the main excitation part but the double-integration was unable
to recover the long-period part reliably, including permanent settlement and rotation because
of the high-pass filtering process.
The resultant forces at the base center point
of the footing were obtained with the bidirectional load cells. Accordingly, the P-Delta
effect, the additional moment caused by selfweight of the structure moving through a lateral
displacement, is automatically included in the
estimated moment, M.

E0 = 24.68 'c0.4776 103 kN/m2

(2)

in which the unit of confining stress, 'c, is


kN/m2.
Fig. 6 shows a schematic diagram of the
model pier footing. The combined V-H-M loadings were achieved through controlling the selfweight of the model pier footing and then
gradually applying a horizontal displacement at
a fixed height via a jack actuator. Displacement
modes and bearing capacities were investigated at different V-H-M combinations by varying the height of the loading point, different soil
densities, and different cyclic loading patterns.
The model pier footing was basically the
same as that used in the shake table experiment. The differences were 1) two cases of pier
height were involved and 2) a universal joint
was attached at the top of the steel rack. Fig. 6
also shows the mass, moment of inertia, and
height of the gravity center from the footing
base in terms of each structural component: a
steel rack, an I beam column, and a footing.
The height of the loading point was varied
using two columns that had different heights.
The same column as that used in the shake table experiment and another taller column having the same I cross-section were employed.
We refer to them as the short column and the
tall column, respectively. The non-dimensional
ratios of applied moment, M, to horizontal force,
H, at the base center point of the footing,
(M/B)/H, became 1.8 and 2.6 (= 1:1.44).
The universal joint was attached to the
model pier foundation to connect with the jack
system. At the universal joint, rotation and up-

CYCLIC LOADING EXPERIMENT PROGRAM

The experiments were conducted at the


Foundation Engineering Laboratory in the Public Works Research Institute, Tsukuba, Japan.
The details in the experiment are reported in a
PWRI report (Fukui et al. 2007a).

278

ward-downward movement were free to occur.


Because of the existence of the universal
joint and the use of different pier heights, the
mass of the model pier footing was different
from that used in the shake table experiment. In
terms of the model pier footing with the tall column, the initial static factors of safety for dead
loads were 28 for a relative density of 80% and
14 for a relative density of 60 %.
The displacement of the model was measured using linear variable displacement transducers (LVDTs) and laser displacement transducers (LDTs). A load cell attached to the actuator measured the load on the actuator.
Slow monotonic lateral loading, one-sided
cyclic lateral loading, and two types of reversed
cyclic lateral loading in the N-S direction were
involved. Fig. 7 shows defined displacement
histories, , where 0 is the reference displacement. We specified the loading patterns, referring to a draft guideline proposed by a joint research of the PWRI and FHWA (Federal Highway Administration in US) for cyclic loading experiments of bridge columns (Unjoh et al. 2006).
The Type I loading pattern was associated with
the Type I seismic motions, while the Type II
loading pattern was associated with the Type II
seismic motions.
The guideline does not include a guideline
on how we determine the reference displacement 0 in the experiment of soil-foundation interactions. In the present experiment, 0 was
assumed based on the result of the preceding
monotonic lateral load case in which the same
experimental conditions except for loading patterns were involved. The displacement level at
which the maximum moment appeared was
taken as the reference displacement 0.
Finally the amplitude of the specific displacement was gradually increased from
0.1250 to 0.250 to 0.50, followed by i 0,
where i = 1, 2, 3, n cycles were repeated at
each displacement amplitude. Some cycles that
had smaller displacement amplitudes than 10
were applied prior to the 10-cycle to investigate
the foundation behavior during small-to-mid
scale earthquakes. As for one-sided cyclic loading, in a loading phase, displacement was controlled and increased up to the specified displacement level. Then, the load was released
down to zero in unloading phases, followed by
the displacement control in the next loading
phase. Lateral cyclic load was planned to apply

279

at an average rate of loading of 10 mm/sec.


However, the actual average rates of loading
were 7.4 mm/sec when the sign of incremental
displacement was positive and 9.7 mm/sec
when the sign of incremental displacement was
negative.
Test series

Test series are tabulated in Table 2, where


``C denotes cyclic loading experiment. We had
9 test cases in total, Cases C1 to C3, Cases C4
to C6, and Cases C7 to C9 were in the same
categories, respectively. When the soil density
was changed from case to case, the soil deposit was fully removed and reconstructed.
Other than that, only the soil of upper 1 m was
removed and reconstructed. In Case C4, first,
monotonic loading was applied, and the experiment was continued involving the one-sided
cyclic loading with the Type II pattern on the
reversed side.
Data processing

Data processing was basically the same in


the shake table experiment. The measured
forces and displacements of the footing are expressed in terms of the displacements and resultant forces at the base center of the footing
as shown in Fig. 4. However, the positive direction of sliding, u, in Fig. 4 coincides with the
South direction in Figs. 5 and 6 in the cyclic
loading experiment.
In the following results, the displacement of
the model was obtained with LVDTs and LDTs.
The resultant forces at the base center point of
the footing were obtained with the bi-directional
load cells arranged as the base of the footing.
VERTICAL LOADING EXPERIMENT PROGRAM

For each soil condition (Dr = 80% and 60 %),


a concentric vertical push experiment was also
conducted. The experiments were conducted in
another test pit in the Foundation Engineering
Laboratory. The test pit had internal dimensions
4 m 4 m in plan and 2 m deep. A stiff steel
box filled with concrete and having a square
base of 0.5 m 0.5 m in plan was used as the
footing. The whole footing base was covered
with pieces of the same sand paper. At an early
loading stage, a partial unloading and reloading
was applied to estimate the elastic (unloading)

ening of a shallow foundation subjected to a


monotonic loading can be expressed in an exponential function. We will also apply the following equation:

rigidity of the soil.


Four LVDTs were set to measure the settlements at the corners of the footing top. The
settlement of the footing was analyzed as the
average measured displacements. A load cell
was attached on the jack. The vertical load at
the footing base was obtained as the sum of the
measured load with the load cell and the self
weight of the footing.
The details of the experiment are also included in the PWRI report of the cyclic loading
experiment (Fukui et al. 2007a).

V / Vm = 1 exp(R0 vpl / Vm)

where Vm is the bearing capacity for concentric


loading and R0 is the initial tangent stiffness.
The fitted work hardening functions are also
shown in Fig. 10, with the values of Vm = 244.8
kN and R0 = 48946 kN/m for a relative density
of 80% and the values of Vm = 128.8 kN and R0
= 26742 kN/m for a relative density of 60%. Fig.
10 shows that Eq. (5) can approximate the work
hardening very well for both dense and medium
dense sand cases.
Fig. 11 shows the moment-rotation relationships and horizontal force-sliding relationships
at the base center of the footing in Cases C1,
C4, and C7. The relationships also can be approximated with exponential curves as shown in
Fig. 11, where only the total displacement was
considered for simplicity.
Therefore, the work hardening of the footing
response to monotonic loading can be modeled
with exponential functions.

TEST RESULTS
Monotonic loading

Fig. 8 shows the results for the concentric


vertical loading experiments. The measured
maximum loads were 244.8 kN for a relative
density of 80% and 128.8 kN for a relative density of 60%.
Fig. 9 shows enlarged views of the partial
unloading-reloading processes. A least mean
square approximation for the unloading paths
provided that the vertical spring constants for
the footing were 89179 kN/m2 for a relative
density of 80% and 76207 kN/m2 for a relative
density of 60%.
A theoretical vertical spring constant that is
based on elasticity and corresponds to a vibration frequency of zero is given as follows (Gazetas 1991):
Kv = 4.54 G (B / 2) / (1 )

Overall behavior in the shake table experiment

Fig. 12 shows the time histories of acceleration at the ground surface, ag, horizontal displacement at the center of gravity of the top
steel rack, a1, and settlement, sliding and rotation at the base center of the footing, v, u, and ,
respectively, in Cases S1-2, S1-4, S1-5, and
S2-2. v and u are normalized with the footing
length, B. The base center of the footing was
uplifted during the excitation. The elongation of
the natural vibration period of the soilfoundation system was observed as the shake
table loading progressed.
The model pier footing was toppled in Case
S1-5. Because the soil surface beneath the
foundation was not recompacted in Case S1-5
before setting the model, Case S1-5 was considered to be more vulnerable.
Comparing the residual vertical displacement after the excitation between Cases S1-2
and S2-2, the residual displacement in Case
S1-2 was much larger than that in Case S2-2,
although the recorded maximum base accelerations at the shake table were similar in both
cases. Fig. 13 shows the model pier footing

(3)

where Kv is the spring constant, G is the shear


modulus of soil, B is the footing length, and is
the Poisson ratio. Using Eq. (3) and assuming
the Poisson ratio is 0.3, the elastic rigidities are
inversed as 55000 kN/m2 for a relative density
of 80% and 47000 kN/m2 for a relative density
of 60%.
If the total vertical displacement, v, can be
decomposed into the elastic component, vel,
and the plastic component, vpl, the plastic component can be calculated as:
vpl = v vel, vel = V / KV

(5)

(4)

The relationships between the load, V, and the


plastic displacement component, vpl, are obtained as shown in Fig. 10. Nova and Montrasio (1991) have suggested that the work hard-

280

states after the excitations of Case S1-2, Case


S1-4, Case S1-5 and Case S2-2. The model
excessively tilted in Case S1-4 and overturned
Case S1-5. Although Case S2-2 was also subjected to the Kobe wave, the residual rotation of
the model was much smaller than those of
Cases S1-4 and S1-5. In Case S2-2, the amplitude of the inputted motion was reduced 80%
than that in Case S1-4 and Case S1-5, and this
can explain why the residual displacement in
Case S2-2 was smaller. Therefore, the accumulation of the displacement was affected by the
number of cycles of earthquake motions as well
as the acceleration intensity.
Photos of the final deformation of the soil
surface after the concentric vertical loading experiment with a dense soil condition and a fullyreversed cyclic loading experiment with a dense
soil condition (Case C8) were shown in Fig. 14.
This indicates that the slip surface spread out
three-dimensionally.
Hysteretic loops

Fig. 15 shows the moment-rotation curves in


Cases C6 and C9. The soil was dense and medium dense, respectively. The areas of loop in
the medium dense soil condition (Case C9)
were somewhat larger than those in the dense
soil condition (Case C6) at the same rotation
amplitudes. It seems that the moment-rotation
relationship is likely to be modeled with a conventional origin and peak-oriented hysteretic
rule.
Fig. 16 superimposes the moment-rotation
curves in Case S2-2 and Case C4, which were
subjected to shake table loading and monotonic
loading, respectively. Dr equaled 80% and the
short column was attached in both cases. In Fig.
16, the negative moment and rotation relationship of Case C4 is shown after reversing the
signs for both moment and rotation. The backbone curves and monotonic loading curve in Fig.
16 agree well. This indicates that hysteretic
curves for any loading conditions can be characterized with the monotonic loading curve.
Examples of the distributions of the reaction
forces normal to the load cells at the foundation
base in Cases S1-2, S2-2, and C6 are shown in
Fig. 17, together with corresponding momentrotation curves. Based on the result of Case C6
shown in Fig. 17(c), the partial uplift initiated at
point b and the corresponding seismic coefficient level (= moment / the height of the center

281

of gravity / self weight) was approximately 0.2.


Fukui et al. (1999) showed that the lateral
seismic intensity coefficients at which the partial
uplift initiated in typical design cases of highway
bridge shallow foundations seemed to range
between 0.07 and 0.25.
As an increase in the rotation, the uplift of
the footing evolved, the effective area of the
footing reduced and the moment reached the
upper limit value. While one-part of the foundation was uplifted, the opposite part continued to
contact with the underlying soil (points d, g, j, s,
and v in Fig. 17). The significant yield of the soil
resistance initiated when the moment attained
the ultimate value. The normal force distribution
at the footing base became a square shape
(points d, j, and v). However, during the unloading phases (points h and k) and the subsequent
reloading phases (points u and x), the effective
contact area did not increase remarkably. This
implies permanent reductions in the footing-soil
contact area. The reduction in the effective contact area resulted in the gradual degradation of
rotational stiffness in unloading paths as indicated in Fig. 15.
Fig. 18 shows the time evolution of the contact area of the foundation base with the soil in
Cases S1-2, S2-2, and C6. In Fig. 18, the
blacken position corresponds to the contacted
part at each time. The gap between the footing
and the soil evolved beneath the footing edges
increased as time elapsed. Finally the alternative foundation rocking made the surface shape
of the soil beneath the footing rounded, as
shown in the photo of Case C6 in Fig. 14(b).
We infer that the vertical load was supported by
sort of a soil-arching mechanism after the excitation.
Fig. 19 shows the settlement -rotation relationship in Cases S1-2, S2-2, and Case C6
(shake table loading with the Type I seismic
motion, shake table loading with the Type II
seismic motion, and fully-reversed cyclic loading with the Type II loading pattern, respectively). The uplift behavior and accumulation of
the settlement of the footing were eminently observed, while the vertical force changed little as
will be shown later. Therefore, we conceive that
there is a coupling between rotational component and vertical component in the foundation
behavior during earthquakes.
Fig. 20 shows the horizontal force-sliding
curve in Case S2-2 (shake table loading). The

ultimate horizontal loads were observed and the


sliding displacement abruptly increased when
the horizontal load reached the ultimate value.
Fig. 21 shows the moment-rotation relationship, the moment-settlement relationship, and
the rotation-settlement relationship at the base
center of the footing in Case C4, in which onesided monotonic loading was applied on the
reversed side following a preceding monotonic
loading. The numbers, 1, 2, 3..., and arrows in
the graphs depict the order of traveling. In the
cycles following the preceding monotonic loading, the reloading path trends toward the previous peak point and seems to be modeled with a
conventional peak-oriented hysteretic rule.
However, the reloading path does not trend toward the point whose absolute displacement is
equivalent to that of the peak point in the preceding monotonic loading phase. Similar characteristic can be seen in the moment-settlement
relationship and settlement-rotation relationship.
Therefore, the characteristics of hysteretic loops
can be described with an origin-oriented and
peak-oriented rule, but it is also likely to be a
function of both positive-side peak rotation point
and negative side-peak rotation point.
Load histories

Fig. 22 shows the time histories of acceleration at the soil surface, ag, acceleration at the
center of gravity of the top steel rack, a1, normalized vertical force at the base center of the
footing, V/V0, normalized horizontal force at the
base center of the footing, H/V0, and normalized
moment at the base center of the footing,
M/V0/(B/2), in Cases S1-2 (Type I motion) and
S2-2 (Type II motion), where V0 is the weight of
each model pier footing and B is the footing
length. The normalized moment is equivalent to
the eccentricity ratio of the vertical load. The
time window in each excitation phase is included in the main excitation part. The variation
in vertical force was small. Sort of flat peak regions were observed in the time histories of a1,
H and M. In Case S2-2, a Type II motion case,
although a flat peak region of M appeared
around a sequential time of 7 seconds, the noticeable elongation of the vibration frequency in
the time history seemed to initiate after a sequential time of 10 seconds. This infers that the
footing reached the bearing capacity at a sequential time before 10 seconds and the upliftrocking behavior gradually prevailed as the

282

yield of the soil beneath the footing edges


evolved. The time history of soil-footing contact
area shown in Fig. 18 backs up this inference.
Nova and Montrasio (1991) suggested a
failure locus in a normalized V-H-M space for
sand as follows:
fcr = h2 + m2 2(1 )2 = 0

(6)

where = V / Vm, h = H / (Vm), m = M /


(BVm), and Vm is the bearing capacity for
concentric vertical loading. , , and are
constants that define the shape of the failure
locus. Although Nova and Montrasio recommended = 0.95, we will assume = 1 hereafter for simplicity. A schematic diagram of the
failure locus is illustrated in Fig. 23. When =
1, the peak values of h and m occur at = 0.5.
is the tangent on the h- envelope at = 0,
while is the tangent on the m- envelope at
= 0. The value of can be assumed to be tan.
Fig. 24 plots the trajectory of observed
forces in the H-M plane and the projection of
the theoretical failure locus, Eq. (6), at V = V0
and with different values of , where the self
weight of the model pier footing was taken into
account as the value of V0 in each case. Fig.
25 shows the trajectory of observed forces in
the V-M plane and the projection of the theoretical failure locus. The values of H at which
the maximum moment appeared were sought
in Cases S1-2, C2, and C8 and used in the
plots of the projections. Figs. 24(a) and 25(a)
are associated with shake table loading of
Cases S1-2, S1-4, S1-5 and S2-2, Figs. 24(b)
and 25(b) are associated with cyclic loading
and dense sand cases, Cases C2, C3, C5 and
C6, and Figs. 24(c) and 25(c) are associated
with cyclic loading and medium dense sand
cases, Cases C8 and C9. In addition, Figs.
24(b) and 25(b) contain the results of both tall
column cases and short column cases. As
Nova and Montrasio (1991) pointed out that
the value of is supposed to be in the range
0.3-0.5, the experimental data agrees with the
theoretical failure envelop with an assumption
of = 0.5 for both dense and medium dense
sands. Figs. 24 and 25 indicate that the difference in the failure locus between the shake
table loading and cyclic loading was indiscernible, when the soil density was identical.

tical displacement during the 20-cycle, 4 0cycle, and 6 0-cycle. The residual settlements
are the settlement when M became zero on the
way from the negative peak moment to the following positive peak moment. While both cases
were involved with Type I cyclic loading pattern,
Case C5 used a dense sand deposit and Case
C8 used a medium dense deposit. The rate of
the accumulation of residual settlement during
the repeated cycles at each specified displacement amplitude did not change as the repeated
cycles increased.
Although an isotropic work hardening rule
worked well in a past research (Nova and Montrasio 1991) for monotonic loads, a kinematic
hardening rule or an isotropic-kinematic hardening rule may be preferred for problems that involve large number of repeated cycles in loads.
Therefore, the examination of the effectiveness
of an isotropic work hardening or other work
hardening theories is highly encouraged to
model the behavior of shallow foundations during earthquakes.
As has been seen above, the uplift of the
footing greatly influenced the foundation behavior such as the shape of hysteretic loop, the
gradual degradation of rotational stiffness, and
the elongation of vibration characteristic. As illustrated in Fig. 29, we assume that the displacement, x, can be decomposed into the elastic component, xel, plastic component, xpl, and
uplift component, xup:

Evolution of cumulative displacement

Fig. 26 shows the relationships between the


residual displacement and the given drift angle
amplitude for Cases C6 and C9 (cyclic loading).
The residual displacement is the displacement
when M equaled zero in transition from the i cycle to the (i + 1)-cycle. The drift angle is defined as the ratio of the displacement at the
point of loading, , to the lever arm length, L, /
L, where we simply disregard the sliding of the
footing when we estimate the given drift angle
amplitude. Fig. 26 also shows the relationships
between the residual rotation and the given drift
angle. Case C6 was conducted on dense sand
and Case C9 was conducted on medium-dense
sand. Other conditions were identical to each
other. The accumulation of the residual displacement in the medium-dense sand case
(Case C9) was faster than that in the dense
sand case (Case C6).
Fig. 27 compares the relationships between
the residual displacement and the given drift
angle amplitude, / L, for Cases C3 and C6.
Case C3 used the tall column, while Case C6
used the short column. Other than that, the experimental conditions were identical. Both
cases showed a similar tendency in the accumulation of the residual displacement, but the
accumulation in Case C6, a short column case,
gradually became larger than that in Case C3, a
tall column case, as the given drift angle increased.
As has been mentioned, in the shake table
experiment, the residual deformation depended
on not only the amplitude of applied base acceleration but the number of cycles in the motion. The residual settlement in Case S1-2 was
much larger than that in Case S2-2, although
applied base accelerations were similar in both
cases. As also shown in Fig. 19(b), in cyclic
loading experiments, the vertical settlement
also increased during the repeated cycles at
constant loading amplitude. Accordingly, based
on cyclic loading experiment results, we scrutinize the characteristic in the accumulation of
residual displacements during the repeated cycles with constant amplitude.
For Cases C5 and C8, we analyze the residual settlement during each repeated cycle at
specified displacement amplitudes of 20, 4 0,
and 6 0. Fig. 28 shows the relationship between the repeated cycles and the residual ver-

x = xel + xpl + xup

(7)

where x is vertical displacement, v, horizontal


displacement, u, or rotation, . From the moment M-rotation relationship of Case C9, a
cyclic loading and medium sand case, we try to
estimate the evolution of each component in
terms of .
At any peak rotation point on the positive
side, we assume that the elastic component, el,
can be estimated with:
el = M / Kr, Kr = 3.6G (B/2)3 / (1 )

(8)

where the value of G / (1 ) is obtained based


on the concentric vertical loading experimental
result via Eq. (3). The plastic component, pl, is
estimated as the rotaion at M = 0 unloaded from
the corresponding peak point. Finally the uplift
component, up, is derived by subtracting the

283

elastic and plastic components, el and pl, from


the total rotation at the corresponding peak
point. Fig. 30 shows the derived -el, -pl, and
-up relationships. The uplift component contributed to the total rotation as much as the
plastic component did, and they were predominant at every total rotation level. On the other
hand, the contribution of the elastic component
was always much smaller than that of the plastic and uplift components.

work hardening rule will be tested via the simulations of the experimental results to find out a
relevant work hardening rule for earthquakeinduced cyclic loading.
REFERENCES
Butterfield, R. and Gottardi, G. (1994) A threedimensional failure envelope for shallow foundations on sand, Gotechnique, Vol. 44, No. 1,
pp. 181-184

CONCLUDING REMARKS

This paper has reported the results of largescale shake table experiments and cyclic loading experiments of model pier footings on sand.
The experiments provided extensive data on
the behavior of shallow foundations subjected
to earthquake loading. The findings that should
be important for the development of numerical
models are summarized as follows:
(1) The residual displacement is dependent on
the number of loading cycles during an excitation.
(2) The backbone curves of hysteretic loaddisplacement loops can be modeled based
on the load-displacement curves for monotonic loading.
(3) The coupling effect among vertical displacement, horizontal displacement, and rotation is needed to be considered.
(4) The uplift significantly affects the foundation
behavior such as the shape of hysteretic loop,
the degradation in the rotational stiffness, and
the elongation of the vibration property. When
we assume that the displacement can be decomposed into elastic component, plastic
component, and uplift component, the uplift
component becomes predominant as much
as the plastic component.
(5) The hysteretic rule should be a function of
both positive and negative moments or both
positive and negative rotations, and a typical
origin and peak-oriented rule is likely to be
relevant in modeling.
(6) The work hardening rule for monotonic loading can be modeled with an exponential function. However, the experimental result also
indicated that a sophisticated work hardening
rule such as a kinematic work hardening rule
may be preferred to an isotropic work hardening rule for repeated loading with constant
amplitude.
It is highly encouraged that several kinds of

284

Cremer, C., Pecker, A., and Davenne, L. (2001)


Cyclic macro-element for soil-structure interaction: material geometrical non-linearities, International Journal for Numerical and Analytical
methods in Geomechanics, Vol. 25, pp. 12571284
di Prisco, C., Nova, R., and Sibilia, A. (2002) Analysis of soil-structure interaction of towers under
cyclic loading, Proc. NUMOG 8, Rome (Pande
& Pietruszczak ed.), Swets & Zeitlinger, Lisse,
pp. 637-642
Faccioli, E., Paolucci, R., and Vivero, G. (2001) Investigation of seismic soil-footing interaction by
large scale cyclic tests and analytical models,
Special Presentation Lecture SPL-05, 4th Int.
Conf. on Recent Advances in Geotechnical
Earthquake Engineering and Soil Dynamics,
San Diego, CA.
Fukui, J., Kimura, Y., Ishida, M., and Kishi, Y. (1999)
An investigation on the response of shallow
foundations to large earthquakes, Technical
memorandum of PWRI, No. 3627, Public Works
Research Institute, Tsukuba, Japan (in Japanese)
Fukui, J., Nakatani, S., Shirato, M., Kouno, T.,
Nonomura, Y., and Asai, R. (2007a) Experimental study on the residual displacement of
shallow foundations during large earthquakes,
Technical memorandum of PWRI, No. 4027,
Public Works Research Institute, Tsukuba, Japan (in Japanese).
Fukui, J., Nakatani, S., Shirato, M., Kouno, T.,
Nonomura, Y., Asai, R., and Saito, T. (2007b)
Large-scale shake table test on the nonlinear
seismic response of shallow foundations during
large earthquakes, Technical memorandum of
PWRI, No. 4028, Public Works Research Institute, Tsukuba, Japan (in Japanese)
Gajan, S., Kutter, B., Phalen, J., Hutchinson, TC.,
and Martin, GR. (2005) Centrifuge modeling of
load-deformation behaviour of rocking shallow
foundations. Soil Dynamics and Earthquake
Engineering, Vol. 25, pp. 773-783

Gazetas, G. (1991) Chapter 15, Foundation vibration, Foundation Engineering Handbook (Fang
ed.), van Nostrand Reinhold, NY.

Nova, R., and Montrasio, L. (1991) Settlement of


shallow foundations on sand, Gotechnique, Vol.
41, No. 2, pp. 243-256

Gottardi, G., and Butterfield, R. (1995) The displacement of a model rigid surface footing on
dense sand under general planar loading, Soils
and Foundations, Vol. 35, No. 3, pp. 71-82

Okamura, M., and Matsuo, O. (2002) A displacement prediction method for retaining walls under
seismic loading, Soils and Foundations, Vol. 42,
No. 1, pp. 131-138

Haya, H., and Nishimura, A. (1998) Proposition of


design method of spread foundation considering
large sale earthquake force, JSCE Journal of
Construction Engineering and Management, No.
595/VI-39, pp. 127-140 (in Japanese).

Paolucci, R. (1997) Simple evaluation of earthquake-induced permanent displacements of


shallow foundations, Journal of Earthquake Engineering, Vol. 1, No. 3, pp. 563-579
Paolucci, R., Shirato, M., and Yilmaz, MT. (2007)
Numerical simulations of shaking table experiments on a shallow foundation test model at
PWRI, Japan, Proceeding of the Second JapanGreece Workshop on Seismic Design, Observation, and Retrofit of Foundation, Tokyo, CD-ROM

Houlsby, GT. and Martin, CM. (1993) Modelling of


the behaviour of foundations of jack-up units on
clay, Predictive Soil Mechanics (Houlsby and
Schofield ed.) Thomas Telford, pp. 339-358
Japan Road Associations. (2002) Specifications for
Highway Bridges, Part IV Substructures, Part V
Seismic design, Maruzen, Tokyo.

Unjoh, S., Hoshikuma, J., and Nishida, H. (2006)


"Draft guidelines for experimental verification of
seismic performance of bridges (Quasi-static cyclic loading tests and shake table tests for bridge
columns)", Technical memorandum of PWRI, No.
4023, Public Works Research Institute, Tsukuba,
Japan

Negro, P., Paolucci, R., Pedretti, S., and Faccioli, E.


(2002) Large-scale soil-structure interaction experiments on sand under cyclic loading, Proc.
12th World Conference on Earthquake Engineering, Auckland, New Zealand, paper # 1191

285

Table 1: Shake table experimental program


Case
identifier
S1-1

Earthquake motions (and soil


preparation)
Sweep wave

Planned maximum
acceleration
50 gal

Observed maximum
acceleration on table
112 gal

Embedded
depth
0 mm

S1-2
S1-3

Shichiho wave
Sweep wave

386 gal
50 gal

601 gal
106 gal

0 mm
0 mm

S1-4

812 gal

712 gal

50 mm

812 gal

726 gal

0 mm

S2-1

(Remove the model, recompact the soil, and settle the


model again) Kobe wave
(Remove the model and settle
it at a different position)
Kobe wave (Remove the
model and soil deposit)
Sweep wave

50 gal

110 gal

10 mm

S2-2

Weakened Kobe wave

650 gal

557 gal

10 mm

S2-3

Sweep wave

50 gal

113 gal

10 mm

S1-5

Table 2: Cyclic loading experimental program


Case
identifier
C1

Loading pattern

Relative soil density

Lever arm length, L

Monotonic loading

80%

1300 mm (Tall column)

C2
C3

One-sided cyclic loading with Type II pattern


Fully-reversed cyclic loading with Type II pattern

80%
80%

1300 mm (Tall column)


1300 mm (Tall column)

C4

80%

900 mm (Short column)

C5

Monotonic loading, followed by one-sided cyclic loading on the reversed-side


Fully-reversed cyclic loading with Type I pattern

80%

900 mm (Short column)

C6

Fully-reversed cyclic loading with Type II pattern

80%

900 mm (Short column)

C7

Monotonic loading

60%

900 mm (Short column)

C8

Fully-reversed cyclic loading with Type I pattern

60%

900 mm (Short column)

C9

Fully-reversed cyclic loading with Type II pattern

60%

900 mm (Short column)

286

Fig 1: Experimental setup for shake table experiments

CL
400
300

500

753
250 22 22
150 300

Steel rack
(Superstructure)
S mG=533 kg
Shake
J=9.184 kg-m2
direction h =0.586 m
G
300
Column
mG=23 kg
J=0.254 kg-m2
hG=0.341 m
400
Footing
200
mG=299 kg
J=8.477 kg-m2
hG=0.131 m
Bi-deirectional
CL
loadcells
Side view Sand paper

500
plan view

Fig 2: Model pier footing used in shake table experiment

287

Acceleration : gal

200
0
200
0

30

60
t : sec
(a) Shichiho wave

90

500
0
500
0

10

20
t : sec

(b) Kobe wave


Fig 3: Input motions

1500

1950
Model

Actuator

CL
Reaction beam
H-200x200
1800
500

3000

2050

CL

900 (Short column case)

Reaction beam
H-200x200
Actuator

Model

Ground

2000

1250 500 1250

W
N

1300 (Tall column case)

Fig 4: Sign convention and notation of load and displacement

3000
1500

Acceleration : gal

400

1700
4000
plan view

Toyoura sand (Dry)


Dr = 80%, 60%
1806

500

1704

Side view

Fig 5: Experimental setup for cyclic loading experiment

288

30

500

Footing
mG = 308 kg
J = 8.744 kg-m2
hG = 0.131 m

900
250 150
300 147
9 22 22

200

400

S
E
Steel rack
Loading direction
(Superstructure)
mG = 566 kg
Y
J = 9.688 kg-m2
400
hG = 0.587 m
300
Column
mG = 23 kg
J = 0.254 kg-m2
hG = 0.341m

200

Loading point

Z
300

E
Steel rack
Loading direction
(Superstructure)
mG = 545 kg
Y
J = 9.367 kg-m2
400
hG = 1.034 m
300
Column
mG = 57 kg
J = 2.904 kg-m2
X
hG = 1.034 m

500

400

Loading direction

Footing X
Bi-direction load cells mG = 301 kg 2
J = 8.537 kg-m
Sandpaper
hG = 0.131 m
500

Bi-directional load cells


Side view
Sandpaper

500

Side view

Plan view

(a) Tall column case

(b) Short column case

: mm

Fully-reversed cyclic loading (Type I)

0.125 0, 0.25 0, 0.5 0 (n=3)


5 0
4 0
3 0
1 0 2 0

n=10 n=10

n=5

n=5

6 0

n=3

n=3

Fully-reversed cyclic loading (Type II)

: mm

0.125 0, 0.25 0, 0.5 0 (n=1)


1 0 2 0
n=3

n=3

3 0

4 0

n=2

n=2

6 0

5 0

n=1

One-sided cyclic loading (Type II)

0.125 0, 0.25 0, 0.5 0 (n=1)


3 0
2 0
1 0
n=2
n=3

7 0

4 0
n=2

n=2

5 0
n=1

n=3

Fig 7: Applied cyclic loading histories

289

8 0

n=2

8 0

n=1

500
Plan view

Fig 6: Model pier footings used in cyclic loading experiment

: mm

250 150
9

1300
300 100
403
22
22

Z
300

Loading point

500

Loading direction

n=1
6 0
n=1

Dr = 80%

Vm

244.8
V : kN

200
128.8
100

Dr = 60%

Vm

0.2

0.4

v /B

Fig 8: Vertical loading experiment results

Dr = 80%

Dr = 60%

V : kN

40

20
Kv

Kv

0
0.002

0.003 0.003 0.004


v /B
Fig 9: Partial unloading-reloading curve in vertical loading experiment results

experimental result(Dr = 80%)


V /Vm =
pl
1 exp((R0v )/Vm) (Dr = 80%)

V: kN

200

100
experimental result (Dr = 60%)
pl

V /Vm = 1 exp((R0v )/Vm) (Dr = 60%)


0

20

40
pl
v : mm
Fig 10: Observed and approximated relationships between vertical force, V, and the plastic component of
vertical displacement, vpl

290

exponential
curve (C1)

exponential
curve (C4)
C1

C7
exponential
curve (C7)

C4

exponential
curve (C7)
exponential
C7
0 curve (C1) C1exponential
curve (C4)
C4

H : kN

M : kNm

2
2
0.05

0.05

: rad
(a) Moment-rotation curve

0
8
u : mm

16

(b) Horizontal force-sliding curve

Fig 11: Observed and approximated moment-rotation and horizontal force-sliding curves: Cases C1, C4, and
C7

0
0.01
0.02

a1 : gal ag : gal

a1 : gal ag : gal

300
500
0
500

v /B

300

v /B

S14, 15

600
0

S22

600

700
0
700
0.02
0
0.02

S15

S14
S22
S22

S15

S14

0.04

u /B

u /B

: rad

0
0.02

0.04
20

30

40

t : sec

50

(a) Case S1-2 (Type I motion)

S15

0.2

: rad

S22
S14

S22
S15

0.3
0

S14
10

t : sec

20

30

(b) Cases S1-4, S1-5, and S2-2 (Type II motion)

Fig 12: Time histories of acceleration at the ground surface, ag, acceleration at the center of the top steel
rack, a1, and settlement, sliding, and rotation of the footing, v, u, and : Cases S1-2, S1-4, S1-5, and S2-2

291

(a) Case S1-2

(b) Case S1-4

(c) Case S1-5

(d) Case S2-2

Fig 13: Model pier foundation states after excitations

(a) Concentric vertical loading (Dense sand)

(b) Case C6 (Type II loading, Dense sand)

Fig 14: Photos of deformation of soil surface after experiments

1.5

C6 (dense
soil)

M : kNm

M : kNm

0
C9
(medium
dense soil)

2
0.1

C4
0
S22

1.5
0.05 0

0.1

0.05 0.1

: rad
: rad
Fig 15: Moment-rotation curves: Cases C6 & C9 Fig 16: Moment-rotation curves: Cases S2-2 & C4

292

1
0.06

200

0.06

0.12

:rad

100
200

0
100
200

1
0

200

100

100

p : kN/m

p : kN/m

100
200

p : kN/m

p : kN/m

p : kN/m

p : kN/m

M : kNm
0

100
200

(a) Case S1-2

200

200

0.02

0
1

0.02

100

0.04

:rad

100
200

p : kN/m

200

100

p : kN/m

100

p : kN/m

p : kN/m

p : kN/m

p : kN/m

M : kNm

0
100
200

100
200

(b) Case S2-2

100

1
b
0

100
200

100
200

0.06

0.12

a
1

p : kN/m

200

p : kN/m

p : kN/m

p : kN/m

M : kNm

0
100
200

(c) Case C6
Fig 17: Distributions of normal forces to the footing base and corresponding moment-rotation curves

293

Shake direction
South
North

50
t : sec

100

Shake direction
South
North

(a) Case S1-2

20
t : sec

40

Shake direction
South
North

(b) Case S2-2

100
t : sec

200

(c) Case C6
Fig 18: Time evolution of the contact area of the footing base with the underlying soil

Upward
displacement
Downward
displacement
0.03

: rad

(a) Case S1-2

5
0
5

Upward
displacement
Downward
displacement
0.08

0.08

: rad

(b) Case C6
Fig 19: Rotation-settlement plots

294

v : mm

v : mm

v : mm

10

Upward
displacement

0
3

Downward
displacement

0.03

0.03

: rad

(c) Case S2-2

H : kN

4
14

14

u : mm
Fig 20: Horizontal force-sliding curve: Case S2-2

0
4

1.6
0.08

0
: rad

2
0

0.08

(a) Moment-rotation curve

4
1.6

1
0
1.6
M : kNm

(b) Moment-settlement curve


Fig 21: Results in Case C4

295

v : mm

12

v : mm

M : kNm

1.6

3
0.08

0
: kNm

0.08

(c) Settlement-rotation curve

ag :gal

0
300
800

V/V0 a1 : gal

ag :gal
V/V0 a1 : gal

300

0
800
1.5
1

500
0
500
600
0
600
1.2
0.6

H/V0

0.3
0
0.3
1

2M/(V0B)

2M/(V0B)

H/V0

0.5

0
1
42

43

44

0.3
0
0.3
1

45

t : sec
(a) Case S1-2

0
1

10

11

t : sec
(b) Case S2-2

12

Fig 22: Time histories of acceleration at the ground, ag, acceleration at the center of the top steel rack, a1, and
normalized vertical force, V / V0, normalized horizontal force, H / V0, and normalized moment, M / V0B / 2, at
the base center of the footing

Fig 23: Typical failure locus for footing under combined load in the V-H-M space

M : kNm

= 0.5

S12

= 0.5

1
0

= 0.5

C2

S22
S14

1
2

C5

C8

= 0.45

C6 C3

= 0.45

C9

= 0.45

S15

0
H : kN

0
H : kN

H : kN

(a) Shake table loading


(b) Cyclic loading on dense sand (c) Cyclic loading on medium dense sand
(S1-2, S1-4, S1-5, S2-2)
(C2, C3, C5, C6)
(C8, C9)
Fig 24: Projection of observed forces and theoretical failure locus in H-M plane

296

= 0.5

S12
S22

= 0.5

C6
C5

C9

S15

2 = 0.45
0

C8
= 0.5

C2
C3

S14

8
V : kN

16

= 0.45

8
V : kN

16

= 0.45

8
V : kN

16

(a) Shake table loading


(b) Cyclic loading on dense sand (c) Cyclic loading on medium dense sand
(S1-2, S1-4, S1-5, S2-2)
(C2, C3, C5, C6)
(C8, C9)
Fig 25: Projection of observed forces and theoretical failure locus in M-V plane

vr : mm

0
10
20

C9 (medium-dense sand)

C6 (dense sand)

C9 (medium-dense sand)

|r| : rad

0.03

C6 (dense sand)
0

0.05

/L

0.1

Fig 26: Relationship between residual displacement and given drift angle amplitude: Cases C6 & C9

vr : mm

0
C3 (tall column)
3
C6 (short column)
0.03
C6 (short column)

|r| : rad

M : kNm

0.02
0.01
0

C3 (tall column)
0

0.05

0.1

/L

Fig 27: Relationship between residual displacement and given drift angle amplitude: Cases C3 & C6

297

vr : mm

10

C5 (dense sand)
C8 (medium-dense sand)
: 20 cycle
: 40 cycle
: 60 cycle

15

20

10

Fig 28: Evolution of residual settlement during the repeated cycles at the 20-cycle, 4 0-cycle and 6 0-cycle:
Cases C5 and C8

el

pl

up

x=x +x +x

pl

up

el

Fig 29: Decomposition of total displacement into elastic, plastic, and uplift components
0.04
el

0.02

pl

el

, ,

up

: rad

:
pl
:
up
:

0
0

0.02

: rad

0.04

0.06

Fig 30: -el, -pl, and -up relationships: Case C9

298

Numerical Modeling of Centrifuge Cyclic Lateral Pile Load


Experiments
N. Gerolymos1, V. Drosos1, S. Escoffier2, G. Gazetas1, J. Garnier2
1

National Technical University of Athens, Greece


Laboratoire Central des Ponts et Chausses, Nantes, France

Abstract
To gain insight into the inelastic behavior of piles, the response of a vertical pile embedded in
dry sand and subjected to cyclic lateral loading was studied experimentally in centrifuge tests
conducted in Laboratoire Central des Ponts et Chausses. Three types of cyclic loading were
applied, two asymmetric and one symmetric with respect to the unloaded pile. An approximately
squareroot variation of soil stiffness with depth was obtained from indirect inflight density
measurements, laboratory tests on reconstituted samples, and well-established empirical
correlations. The tests are simulated using a cyclic nonlinear Winkler spring model which can
describe the full range of inelastic phenomena, including separation and re-attachment of the
pile from and to the soil. The model consists of three mathematical expressions capable of
reproducing a wide variety of monotonic and cyclic experimental py curves. The physical
meaning of key model parameters is graphically explained and related to soil behavior.
Comparisons with the centrifuge test results demonstrate the general validity of the model and
its capability of capturing several features of pilesoil interaction, including : soil plastification at
an early stage of loading, pinching behavior due to the formation of a relaxation zone around
the upper part of the pile, stiffness and strength changes due to cyclic loading. A comparison is
presented of the py curves derived from the test results with those obtained with the proposed
model, as well as with the classical curves of Reese (1974) for sand.

INTRODUCTION: EMPIRICAL AND ANALYTICAL


py CURVES FOR LATERALLY LOADED
PILES

The response of piles subjected to cyclic


lateral loading is governed by the strong
nonlinearity of the stressstrain soil behavior
occurring even at low levels of the applied load.
The problem becomes much more difficult with
the appearance of geometric nonlinearities,
such as separation of and sliding between pile
and soilphenomena unavoidable under
strong excitation.
The methods of lateral pile response
analysis could be classified into three broad
categories :

hinging has transformed the pile into a


mechanism (Hansen 1961, and Broms 1964
(a,b)).
elastic (and inelastic) continuumbased
methods which rarely lead to analytical
solutions, but are usually materialized
through boundaryelement, finiteelement,
or
finitedifference
type
numerical
formulations (see Banerjee (1978a,b),
Poulos & Davis 1980)
linear and nonlinear Winkler spring
methods, the most successful of which is
the py method (McClelland & Focht, 1958 ;
Matlock 1970 ; and Reese et al 1974, 1975,
1986)

A finite element analysis requires discretization


of the pile and the surrounding soil in 3
dimensions. Equivalentlinear as well as
advanced constitutive models based on the
theory of plasticity and hypoplasticity have been

limit analysis methods in which the ultimate


soil reaction is predetermined from the
assumed shape of the pile displacement
profile at its ultimate state, i.e., after plastic

299

components in series, and which can thus


model gapping effects, strength degradation,
and radiation damping (Boulanger et al, 1999).
Similar models were presented by Nogami et al
(1992).
Several
other
approximate
nonlinear
methods, of a different philoshophy, have also
been developed over the years. They are of
significant engineering interest, but are beyond
the scope of this paper. Just for completeness
of our discussion we refer only to Duncan et al
(1994), Budhu & Davies (1987) and
Kucukarslan & Banerjee (2004) for monotonic
loading conditions, and to Tabesh & Poulos
(2001) for seismic response, among several
other publications. A comprehensive review on
the subject was compiled by Pender (1993).

utilized to reproduce the nonlinear stressstrain


soil behavior (examples : Angelides & Roesset,
1981; Trochanis et al., 1991; Kimura et al.,
1995; Wakai et al., 1999). However, even
today, a (3-D) finite element analysis is not a
computationally trivial task and is thus used in
piling practice rather infrequently. Modeling
pilesoil separation and gap formation as well
as other interface nonlinearities can prove
especially formidable tasks. Additionally, a 3D
elastoplastic finite-element model would not be
easily combined with many structural codes to
compute the response of the superstructure.
In contrast, Winkler-spring modeling is a
versatile and economical approach since the
analysis of soilpile interaction is effectively
reduced to a onedimensional problem. It is in
essence a semiempirical method, in which soil
resistance is represented by independent
springs distributed along the pile. The Winklerspring model for laterally-loaded piles owes its
popularity to the well known semi-empirically
derived py curves. Obtained on the basis of
fullscale experiments, they relate soil reaction
(the horizontal resultant of the soil stresses)
with pile deflection at each point of the pile. The
main advantages of the py method is that it
can easily accommodate other experimental
results, as necessary; it (indirectly) accounts for
soil separation from the pile and sliding at the
pilesoil interface; and it can even account for
the method of pile installation.
The
development of py curves has been addressed
by several researchers (e.g., Matlock 1970;
Reese et al., 1974, 1986; Yegian & Wright,
1973 ; Stevens & Audibert, 1979 ; O Neill &
Murchinson 1983; Murchinson & O Neill 1984;
Wu et al., 1998; Ashour & Norris 2000, Kim et
al 2004), and the py method constitutes the
current state of the art . Its success stems from
the fact that even though the py curves do not
model accurately the soil continuum, they are
based on results of field load tests where the
continuum is fully satisfied (Reese 1997).
It is worth mentioning that numerous (mostly
successful) attempts have also been published
developing cyclic py curves for different soils :
starting from the conceptual framework of
Matlock et al (1978), to the development of the
composite
centrifugebased
py
macroelement of Curras et al (1999, 2000),
which consists of viscoelastic, plastic and gap

Most of the aforementioned py methods


employ a semiempirical approach to
developing py curves : essentially the
proposed curves are a judicious curve fitting to
appropriate
fullscale
(or
centrifuge)
experiments. (Judicious meaning with the help
of soil mechanics and engineering judgment.)
A different, semitheoretical methodology to
developing cyclic py curves is also possible.
One starts with a mathematical model (a py
macroelement) and then calibrates its
parameters with the help of fullscale and
centrifuge experiments, or even with rigorous
3D numerical results, if such are available.
Validation of the resulting method against other
test results would then be necessary before
such a method is adopted in practice.
Among a number of such mathematical
models proposed over the last twenty years,
particularly fruitful has proved the socalled
BoucWen model (Bouc 1971, Wen 1976).
Originally applied to describe inelastic cyclic
forcedisplacement relationships in probabilistic
structural dynamics (Baber & Wen 1981), it was
subsequently applied to soil liquefaction
analysis as a constitutive relationship in
simple shear (Pires 1989, Loh et al 1995) and
to the analysis of laterally loaded piles as a
monotonic or cyclic py relationship (Trochanis
et al 1994, Badoni & Makris 1996). The latter
references are of particular interest in the work
presented here, as they showed that a BoucWen model is capable of describing in sufficient
engineering detail the response of a pile under

300

cyclic and dynamic loading at least in cases


of simple soil conditions.
Recently, Gerolymos and Gazetas (2005 a,
b) presented an extensionmodification of the
original BoucWen model, and they applied it to
describe dynamic simple shear relations for
wavepropagation siteresponse analyses
(2005 a) and py relationships for laterally
loaded piles (2005 b). Designated as BWGG
model, this extended BoucWen model is
capable of reproducing complex features of
pilesoil interaction, such as : (i) soil and pile
nonlinearites
;
(ii)
soilpile
interface
nonlinearities ; (iii) coupling between radiation
damping and hysteretic soil response, and (iv)
stiffness
and
strength
hardening
(or
degradation) with cyclic loading. The model was
validated against available experimental data.
The
need,
however,
for
further
calibration/validation
of
the
model
is
unquestionable hence the work presented in
this paper.
The objectivestasks of this paper are : (i) to
briefly introduce the key features and some
capabilities of the developed BWGG model ; (ii)
to present the results of cyclic centrifuge
experiments on a pile in sand ; (iii) to outline the
methodology for calibrating the model
parameters using the results of one of the tests
; (iv) to apply the method to the other tests and
to compare with the centrifuge experiments as
well as with the classical py curves of Reese
(1974). It is hoped that, in the process, some
valuable insight into the nature of pile lateral
response will be gained.
Your goal is to simulate, as closely as
possible, the format of this page. Authors
affiliations should appear immediately following
their names. Sub-sections are not numbered.

Centrifuge Modeling
It is well known that since soils have stressdependent stiffness and strength characteristics
the application of N times the gravitational
acceleration to a model with length dimensions
1/N, makes the stresses and mechanical
properties of the model to become similar to
those of the prototype. Schofield (1980, 1981)
and Whitman et al (1981) among several others
have discussed the scaling laws which relate
the behaviour of a model under static and
seismic shaking to the prototype behavior in the
field. Over the past decade, dynamic centrifuge
techniques have been established as a useful
tool for the engineer to investigate the dynamic
behavior of geotechnical structures and to
calibrate advanced numerical models and
procedures.
The centrifuge facility of the LCPC, in
Nantes, has a radius of 5.5 m, a maximum
model mass of 2000 kg at a centrifuge
acceleration of 100 g, and platform dimensions
of 1.4 m x 1.15 m. It is currently capable of
producing 200 gs of centrifugal acceleration,
although of course at a much reduced
payload. The recent acquisition of a servohydraulic earthquake actuator has extended the
scope of its activities.
Model Description
The presented cyclic lateral load tests were
conducted on vertical friction pile placed in a
sand mass of uniform density. The
Fontainebleau sand centrifuge specimens
were prepared by the air sandraining process
into a rectangular container (80 cm wide by 120
cm long by 36 cm deep), with the use of a
special automatic hopper developed at LCPC
(Garnier, 2002). The desired density of the dry
sand was obtained by varying three
parameters: (a) the flow of sand (opening of the
hopper), (b) the automatically maintained drop
height, and (c) the scanning rate. The unit
weight and the relative density of the specimen
were measured to be d 16.5 0.04 kN/m3
and Dr = 86% samples, respectively. Laboratory
results from (drained and undrained) torsional
and direct shear tests on Fontainebleau sand
reconstituted specimens indicated mean values
of peak and criticalstate angles of p = 41.8

CENTRIFUGE EXPERIMENTS

The centrifuge tests reported here were


conducted for the dissertation of Rosquot
(2004) at Laboratoire Central des Ponts et
Chausses [LCPC] see also Rosquot et al
2003, 2004. The tests were performed on a
single pile subjected to cyclic horizontal loading.
The centrifuge models, 1/40 in scale, involved
pile head loading with three different force time
histories. The experimental set up and the
loading time histories (in prototype scale) are
portrayed in Fig 1.

301

Disp. Sensor DT2


Calculated dis. DPC

1m
Load

Disp. Sensor DT1

20 pairs of strain gauges

1.6 m

Fontainebleau

Pile

Sand

12 m
0.6 m

d = 0.72 m

960

Load : kN

1000

480

500
0
-500
-1000

P32 test
0

10

15

20

25

30

35

40

960

Load : kN

1000
500

0
-500

P344 test
-1000

10

15

20

25

30

35

40

P330 test

1000

Load : kN

960

500
0
-500

960
-1000

10

15

20

25

30

35

40

time : s

Fig. 1: (a) Experimental setup of the centrifuge tests conducted in LCPC. (b) Load time histories of the three
tests (P32, P344 and P330). All dimensions refer to the modeled prototype.

302

and cv = 33, respectively. The reader is


referred to the official site of the Quaker (2002)
research program for details on the
aforementioned tests.
The utilized soil properties of the centrifuge
tests are depicted in Fig 2, with emphasis on a
probable but idealized profile of the shear
modulus Go at small strains. The accuracy of
this profile does not have a substantial effect on
our numerical results since we utilize one of the
tests for calibration of the model. Evidently, in
such a relatively-dense sand, the pile used may
be considered as flexible (Rosquoet 2004;
Randolph 1981).

differentiation of M(z) as established by Matlock


and Reese (see for instance Reese & Van Impe
2001). The strain gauges were spaced at 0.6 m
(prototype scale) starting from the ground level
to the pile tip. This single pile was driven into
the sand before rotating of the centrifuge (i.e.,
at 1 g). In flight, the single pile was subjected
quasi-statically to horizontal cyclic loading
through a servo-jack connected to the pile with
a cable. With such a configuration the pile head
is not submitted to any parasitic bending
moment.
Three cyclic load tests were performed as
sketched in Fig. 1 and discussed later. The test
results were obtained in the form of horizontal
forcedisplacement time histories at the head of
the pile, as well as of bending moment, shear
force, and soil reaction profiles.

0.72 m
1m

G0 : MPa

2.6 m
0

z:m

40

80

120

DESCRIPTION OF THE THEORETICAL MODEL:


QUATIONS AND PARAMETERS
G0 ( z ) 30 z

The developed BWGG model is a versatile


onedimensional actionreaction relationship,
capable of reproducing an almost endless
variety of stressstrain or forcedisplacement or
momentrotation relations, monotonic as well
as cyclic. It is being applied here to model the
monotonic and cyclic response of piles,
expressing the py relationship. A simple
version of the model is outlined below. More
details can be found in Gerolymos & Gazetas
(2005 a, b), although the model utilized here is
a slightly improvedsimplified version of the
model in the latter reference.
The constitutive relationship for the lateral
soil reaction against a deflecting pile is
expressed as the sum of an elastic and a
hysteretic component according to:

4
12 m
8

12
41.8o
d 1.63 t/m3

Fig. 2: Pile configuration and soil properties of the


three centrifuge tests

The model pile (scale 1/40) is aluminium


hollow cylinder of 18 mm external diameter, 3
mm wall thickness, and 365 mm length. The
flexural stiffness of the pile is 0.197 kN m2 and
the elastic limit stress of the aluminium is 245
kPa. The model and prototype pile
characteristics are given in Table 1 (centrifuge
tests were carried out at 40g).
The
instrumentation
included
two
displacement sensors, located at the section of
the pile above the ground line, and 20 pairs of
strain gauges, positioned along the length of
the pile so that the bending moment profile M(z)
could be measured during the tests. The
resultant earth pressure p = p(z), per unit length
along the pile, was obtained by double

p = k y + (1 ) p y

(1)

where is a dimensionless inelastic parameter


expressed in the following differential form :

d
1
=
1
dy
y0

[b + g sign(dy )]}

(2)

p is the resultant (in the direction of loading) of


the normal and shear stresses along the
perimeter of a pile segment of unit length ; y is

303

soil reaction ; and


n, b and g, are
dimensionless quantities that control the shape
of the hysteretic soil reactionpile deflection
loop as describe below.
Parameter n governs the sharpness of the
transition from the linear to the nonlinear range
during initial virgin (monotonic) loading (Fig. 5).
It can take values between 0 and . A large
value of n (> 10) models approximately a
bilinear hysteretic curve ; decreasing n leads to
smoother transitions, with plastic behaviour
occurring at lower loading levels. Parameter
is the ratio of steadystate post yielding to the
initial elastic stiffness. The larger the parameter
, the larger the component of the lateral soil
reaction resulting from constrained soil
dilatancy. These two parameters, n and ,
properly calibrated we can approximately match
: (a) most lateral py curves, such as those
proposed by Reese (1974, 1975) and Matlock
(1970), and (b) almost any experimental soil
reaction curve. Figs 3 and 4 illustrate the
significance of a and n on the resulting py
curves.

Table 1: Pile Characteristics


Name

Symbol

Model
scale

Prototype scale
(40g)

Length

38 cm

15.2 m

Depth of pile tip


from ground
surface

30 cm

12 m

External
diameter

1.8 cm

0.72 m

Internal diameter

1.5 cm

0.6 m
4

Youngs modulus

7.4 x 10 MPa

Moment of
inertia

Bending stiffness

EI

Elastic limit

2.67 x 10
4
m

-9

-3

6.83 10 m

197 N m

505 MNm

245 MPa

1.4

Normalized Soil Reaction ( p / py )

Normalized Soil Reaction ( p / py )

the pile deflection at the location of the spring;


k is a reference spring stiffness ; is a
parameter governing the post yielding stiffness ;
py is a characteristic value of the soil reaction
related to the initiation of significant inelasticity
(yielding) ; y0 is a characteristic value of pile
deflection related to the initiation of yielding in

1.2
1

10

0.8

0.5
0.2

0.6

n = 0.1

0.4

=0

0.2
0
0

Normalized Pile Deflection ( y / y0 )

1.4
1.2
1

10

0.5

0.8

0.2

n=

0.1

0.6
0.4

= 0.05

0.2
0
0

Normalized Pile Deflection ( y / y0 )

Fig. 3: Normalized soil reaction (p / py) versus pile deflection (y / y0) curves in monotonic loading for selected
values of parameter n , derived from the proposed model for = 0 and 0.05

nonlinear but near elastic behavior. In the


special case where b = 1 and g = 0 , the
hysteretic loop degenerates to the monotonic
loading curve (nonlinear but elastic
behavior). On the contrary, as g tends to 1,
the reversal stiffness becomes larger than the
initial stiffness (at virgin loading). When g = 1
and b = 0 the reversal stiffness is two times
larger
than
the
initial
one.
The
unloadingreloading parameters b and g can
be either constants or variable in the course of

Parameters b and g control the


unloadingreloading rule. Four basic shapes
of hysteresis loops can be generated
depending on the relation between b and g
(Fig. 5). When b = g = 0.5, the stiffness upon
reversal equals the initial (maximum) stiffness
and,
the
Masing
criterion
for
loadinguploadingreloading is satisfied. As b
tends to 1, the reversal stiffness tends to be
equal to the post yielding stiffness right before
the reversal point, which implies a strongly

304

resistance under cyclic loading.

1.4
15
0.

1.2

0.1

Normalized Soil Reaction ( p / py )

Normalized Soil Reaction ( p / py )

cyclic loading. The condition b + g = 1 along


with a = 0
imply a constant ultimate

5
0.0

=0

0.8
0.6
0.4

n = 0.2

0.2
0
0

Normalized Pile Deflection ( y / y0 )

1.4
15
0.

1.2

0.1

0.05

=0

1
0.8
0.6
0.4

n=1

0.2
0
0

Normalized Pile Deflection ( y / y0 )

Fig. 4: Normalized soil reaction (p / py) versus pile deflection (y / y0) curves in monotonic loading for selected
values of postyielding parameter , derived from the proposed model for n = 0.2 and 1

to the contact region of the lateral soil reaction.


The continuous nature of the pinching
function produces smooth hysteresis loops with
gradual transition from almost zero stiffness at
complete separation just when to the maximum
stiffness when re-attachment occurs. Parameter
in Eqn (3) controls the gap growth during
cyclic (repeated) loading of the pile. Parameter
o, determines the drag (i.e. side shear)
resistance within the gap. It takes values
between 0 and 1, with the drag resistance
becoming negligible (sharp separation) when 0
approaches 1. In case of a stiff cohesive soil a
value of 0 1 could be adopted if loading in
two direction generates a complete gap around
the pile (sharp separation, negligible drag
forces). On the other hand, when the pile is
embedded in a cohesionless soil, the formation
of a relaxation zone around the pile is the most
possible local failure mechanism near the
ground surface (diffuse failure), rather than the
opening of a clear gap (as in a cohesive soil). In
that case, values of 0 < 1 close to 0 should be
used.
Parameter = (y) is the maximum
attained displacement when y is positive, or the
minimum attained displacement when y is
negative.
A simplified but reasonable hypothesis is
that separation takes place when the net tensile
stress at a point of pilesoil interface becomes

From Equations (1) and (2) the parameter


can be eliminated through a straightforward
stepbystep numerical integration, which is
readily implemented within the framework of
codes such as Mathcad and Matlab.
Modeling Seperation / Gapping of the Pile
from the Soil
A significant issue in pilesoil interaction
under cyclic lateral loading is the separation
(gapping) of the pile from the soil and the
subsequent re-attachment. Separation at the
pilesoil interface is implemented in the
proposed constitutive law for lateral soil
reaction, through the introduction in the
differential expression for (Eqn (2)) of a
multiplicating pinching function of , hp:

y 2
h p ( ) = 1 0 exp 0
( y )

(3)

such that Eqn (2) becomes :

h p ( )
d
=
1
dy
y0

[b + g sign(dy )]}
(4)

In Eqn (9) 0, , and are dimensionless


parameters that control the gap growth and the
sharpness of the transition from the nocontact

305

of situations in which gapping dominates.

larger than the product of the difference


between the earth pressure at rest ho and the
lateral active earth pressure ha , multiplied by
the pile diameter d, i.e. :

'
h0

'
ha
d < p

Stiffness and Strength Parameters for


Cohesionless Soil
The
following
expression
for
the
smallamplitude stiffness k (= py / y0 ) in Eqn(1)
is adequate:

(5)

2
1.5

b = 0, g = 1

1
0.5
0
-0.5
-1
-1.5
-2
-6

-4

-2

k = 1.2 E s

Normalized Soil Reaction ( p / py )

Normalized Soil Reaction ( p / py )

Fig. 6 portrays lateral soil reaction versus


pile deflection hysteresis loops with gapping
effect, for different values of parameters 0 and
. Evidently, the model can reproduce a variety
2
1.5

0.5
0
-0.5
-1
-1.5
-2
-6

1
0.5
0
-0.5
-1
-1.5
-2

Normalized Soil Reaction ( p / py )

Normalized Soil Reaction ( p / py )

b = 0.9, g = 0.1

-4

-4

-2

Normalized Pile Deflection ( y / y0 )

-2
-6

b = 0.5, g = 0.5

Normalized Pile Deflection ( y / y0 )

1.5

(6)

2
1.5

b = 1, g = 0

1
0.5
0
-0.5
-1
-1.5

Normalized Pile Deflection ( y / y0 )

-2
-6

-4

-2

Normalized Pile Deflection ( y / y0 )

Fig 5: Normalized hysteresis loops of soil reaction versus pile deflection for n = 1 and different values of b and
g. The Masing criterion for unloadingreloading is recovered for b =0.5, g = 0.5.

could also be used in other cases, such as


those by Matlock and Reese (in the aforesaid
publications) from the scaled field tests.
In case of a pile embedded in cohesionless
soil the strength parameter py in Eqn (1) can be
calculated through the analytical expression of
Broms (1964) for the ultimate soil reaction:

where Es is the Youngs modulus of the soil


(e.g., Gazetas & Dobry (1984). Note that k has
units of stiffness per unit length of the pile,
corresponding to the traditional subgrade
modulus (in units of pressure per unit length)
multiplied by the diameter d of the pile. Eqn (6)
has been derived by matching the dynamic
head displacement of the Winkler model with
the one computed through finite element
analyses of a soil with Poissons ratio = 0.4.
Other expressions for the subgrade modulus

p y = 3 ' tan 2 45 + d
2

306

(7)

M(z), obtained from the bending strains


measured during each test through the strain
gauges, were utilized to calculate the shear
force, Q(z), and soil reaction, p(z), diagram :

in which v is the effective vertical stress and


is the friction angle of the soil. Eqn (7) is very
often preferred in practice from other more
rigorous expressions for its simplicity and
sufficient engineering accuracy, in view of its
accord with experimental and numerical results.
It is used in all the subsequent analyses as
well.

2 y( z, t )
E p I p
+ p( z, t ) = 0
z 2

d M ( z)
dz

p( z ) =

d 2 M ( z)
dz 2

(9a)

and

Numerical Formulation for PileSoil System


With the constitutive model for lateral soil
reaction developed in the previous sections, the
pilesoil interaction problem under cyclic lateral
loading reduces to the analysis of a beam
supported on a nonlinear Winkler foundation.
Equilibrium of the pile gives :

2
z 2

Q( z ) =

(9b)

High-order spline functions which interpolate


between two successive pairs of experimental
points (Mi, zi) and (Mi+1, zi+1) were utilized to this
end. The experimental M = M(z)curves were
also integrated twice to get the pile deflection
diagram y = y(z) and the boundary conditions
(head and tip pile displacement) were used to
determine the needed two constants.

(8)

where Ep and Ip are the pile modulus of


elasticity and crosssectional moment of inertia,
respectively.
Eqn (9) forms a system of
(coupled) ordinary and differential equations
with Eqns (1), (3), and (4). An explicit
finitedifference scheme is used for the solution
of this system, with the possibility of considering
the variation of pile and soil properties along the
pile length. Head and tip boundary conditions
are properly taken into account.
The proposed model for piles is next applied
to simulate the pile response in the centrifuge
cyclic lateral load tests described in the
beginning.

Calibration of Model Parameters against


Test P32 ; Comparison with empirical py
curves
The calibration of the model parameters was
based on matching the calculated with the
recorded forcedisplacement curve at the pile
head (calculated displacement DPC in Fig 1).
Only density measurements were performed
Hardins (1978) formula is applied to evaluate
the soil shear modulus at lowamplitude strains:

G0 =

NUMERICAL SIMULATION OF THE


CENTRIFUGE TESTS

The proposed model is first calibrated


against the results of one of the tests
(designated as P32). Subsequently it is applied
to predict the measured data of the other two
tests (P330 and P344). The three tests differ by
the characteristics of the cyclic loading
sequences as shown in Fig 2 (one-way and
two-way loading at different load amplitudes). It
should be noted that the applied loads always
stay in the domain of service loads. The
maximum applied load of 960 kN is indeed just
about one third of the maximum lateral
resistance of the pile (Rosquoet, 2004 ; Broms
1964).
The bending moment distribution with depth

(1 + 2 K 0 ) p a v
3 (1 + ) (0.3 + 0.7 e02 )

S
2

(10)

in which the stiffness coefficient S varies in the


range of 1200 and 1500 for clean sands, K0 is
the coefficient of earth pressure at rest, e0 is the
initial void ratio, and pa is the atmospheric
pressure. For e0 0.59, and assuming K0 = 0.5,
= 0.4, and S = 1400, the maximum shear
modulus at the effective stress level of 0.1 MPa
is approximately equal to 75 MPa. The
distribution of G0 with depth is depicted in Fig 2.
The initial subgrade modulus of the nonlinear
Winkler springs is calculated from Eqn (6),
whereas Broms formula [Eqn (7)] is utilized to
estimate the ultimate soil reaction. The
unloadingreloading parameter b is taken equal
to 0.50.

307

Normalized Soil Reaction ( p / py )


Normalized Soil Reaction ( p / py )

Normalized Soil Reaction ( p / py )

1.5
0 = 0.99, = 0.054

=0

0.5
0
-0.5
-1
-1.5
-15
1.5

-10

-5

10

15

10

15

10

15

Normalized
Pile Deflection ( y / y0 )
0 = 0.99,
= 0.054
= 0.03

0.5
0
-0.5
-1
-1.5
-15
1.5
1

-10

-5

Normalized
Pile Deflection ( y / y0 )
= 0.054
0 = 0.97,

= 0.03

0.5
0
-0.5
-1
-1.5
-15

-10

-5

Normalized Pile Deflection ( y / y0 )


Fig. 6: Normalized hysteretis loops of soil reaction versus pile deflection, including separation of the pile from
the soil derived from Eqns [2], [3], and [6] : (a) No drag force and no strength hardening develop
during separation [ n = 1, b = 0.5, = 0, 0 = 0.99, and = 0.054]; (b) Substantial strength hardening
but no drag force develop during separation [n = 1, b = 0.5, = 0.03, 0 = 0.99, and = 0.054]; and (c)
drag force of the order of 10% of the ultimate soil reaction and significant strength hardening develop
during separation [ n = 1, b = 0.5, = 0.03, 0 = 0.97, and = 0.054]

308

Soil Reaction p : kN / m

1500

6m

6m

3.6 m

3.6 m

1.8 m

1000

500

1.8 m

0.02

0.04
0.06
Displacement y : m

0.08

0.1

Soil Reaction p : kN / m

1500

6m
1000

.3
n=0

3.6 m
.25
n=0

500

1.8 m
n = 0.2
0

0.02

0.04

0.06

0.08

0.1

Displacement y : m
Fig. 7: (a) Comparison of the p-y curves at three different depths derived with the proposed model (solid lines)
after calibration with test P32, and those developed by Reese & Matlock for sand (dashed lines).
Based on best fitting the recorded forcedisplacement curve at sensor DPC. (Derived model
parameter values: n = 0.05, = 0.03 ). (b) Comparison of the py curves at the same three depths
derived with the proposed model (solid lines) by fitting Reeses (1974) curves. (Derived model
parameter values: n = 0.20.3, = 0).

monotonic loading ; and (b) n = 0.15 and =


0.025 for cyclic loading. The pinching
parameters were calculated as 0 = 0.95 and
= 0.054. Note that the parameter n is not kept
constant : its initial value of 0.05 changes to

A trialanderror method was used to adjust


the values of the pinching parameters 0 and ,
and the monotonic loading parameters n and .
The following combination was found to give
the best results : (a) n = 0.05 and = 0.03 for

309

bending moment, shear force, and soil reaction


distributions with depth, is presented in Figs 12
and 13 for various stages of loading
The agreement between measurements and
computations is in general quite satisfactory for
the forcedisplacement curve at the pile head,
as well as for the bending moment and shear
force diagrams with depth. The model is
capable of reproducing :
(a) the highly nonlinear soil behaviour even
at low loading levels,
(b) the stiffer reversal stiffness of the pile
response compared to the virgin loading
(initial) stiffness, and
(c) the strength relaxation of the pile
response with cyclic loading. The
displacement at which the maximum
applied external force occurs, increases
with increasing cycle number .
The shapes of hysteresis loops, although
not in such a good agreement with the
experimental curves, reveal the beginning of a
pinching behavior apparently the result of
pilesoil separation and re-attachment near the
top.
Some differences are also noted between
computed and measured soil reaction profiles.
More specifically, the maximum soil reaction is
underestimated by the proposed model,
whereas soil yielding at greater depths is
overestimated. This discrepancy may be
attributed to several factors, including: (a) the
approximate nature of the Go profile which, after
all , was not obtained from direct in-flight
measurements, and (b) the conservatism in
Broms formula for ultimate soil reaction (Eqn
7). Nevertheless, it is rather satisfactory that
the model captures the downward migration of
the maximum soil reaction with increasing
loading intensity apparently the result of
progressively increasing depth of soil yielding
near the top.

0.15 after the first reversal in the loading


history, reflecting the stiffness hardening of the
pile response under cyclic loading due to
densification of the sand.
Fig (7a) compares the py curves at depths
of 6 m, 3.6 m, and 1.8 m computed with the
proposed model to those of Reese (1974) for
the particular sand. The calculation of the py
curves is based on the best fitting of : the P32
test , and (b) Reeses curves. In order for the
py curves to be in a comparable form, Eqn (6)
for the initial subgrade modulus of the spring
and Eqn (7) for the ultimate lateral soil reaction,
have also been implemented for calculation of
slightly modified Reese curves. It is interesting
to note that the inelastic component of the
lateral soil reaction is underestimated with
Reeses curves, while the initial stiffness is
overestimated.
Different py curves would
result from the proposed model if a direct fitting
onto Reeses py curves is performed. Such
fitting is shown in Fig 7(b), only for comparison
purposes.
Fig 8 shows the comparison between
measured and computed (from the calibrated
model) forcedisplacement curve at the head of
the pile. The good agreement attests only, the
capability of our model to reproduce the cyclic
py curves. Figs 9 and 10 compare computed
versus recorded bending moment, shear force,
and soil reaction profiles at different stages of
loading. The success of this comparison is an
indication of proper calibration of the model
against this centrifuge test ; it is encouraging
only in view of the well known difficulty to match
simultaneously detailed forcedisplacement
time history, internal pile forces, and soil
reaction profiles.
Comparison of Numerical Results with
Two Cyclic Experiments
The calibrated model (from the P32 test) is
utilized to predict the results of the P344 test
(cyclic loading without sign reversal) and the
P330 test (fully cyclic loading with sign
reversal).

Test P330 : Symmetric Cyclic Force


between 960 kN

Ranging

Similarly satisfactory are the comparisons


between theoretical predictions and the results
of the P330 test presented in Figs 14, 15 and
16. Specifically, it is seen that the model
predicts well the following detailed trends:

Test P344 : Asymmetric Cyclic Force Ranging


from 960 to 0 kN

The computed forcedisplacement curve at


the pile head is compared to the experimental
data from the P344 test, in Fig 11. The
comparison of the predicted to the measured

(a) the stiffness hardening of the pile under


cyclic loading, possibly due to densification
of the sand in the vicinity of the pile. (Notice

310

Notice also that the maximum bending


moment is slightly underpredicted, whereas the
soil reaction with depth is better captured than
in P344 test.

that the secant stiffness increases slightly


with increasing number of cycles.)
(b) the narrowing of the forcedisplacement
hysteretic loop which could possibly be
attributed to the beginning of formation of a
relaxation zone (gapping) around the pile
and near the ground surface.

1000

Force at Pile Head : kN

Experimental
800

600

400

200

P32 test
0

0.04

0.08

0.12

0.16

0.2

0.16

0.2

Displacement : m
1000

Force at Pile Head : kN

Calculated
800

600

400

200

P32 test

0.04

0.08

0.12

Displacement : m

Fig 8: (a) Experimental and (b) computed total forcedisplacement curve at sensor DPC for test P32 which
was utilized in the calibration

311

Bending Moment : kNm


-500
0

1000

2000

3000

Bending Moment : kNm


4000

-500
0

1000

244 kN

2000

3000

4000

at 957 kN

548 kN

12th cycle
Depth : m

at 957 kN
6

10

10

12

12

Fig. 9: Comparison of computed (solid lines) and recorded (circles and triangles) bending moment
distributions at different stages of loading, for test P32 which was utilized in the calibration

Soil Reaction : kN / m

Shear Force : kN
-1200
0

-800

-400

400

800

-200
0

548 kN
at 957 kN

Depth : m

200

400

600

244 kN
548 kN
at 957 kN

244 kN
6

10

10

12

12

Fig. 10: Comparison of computed (solid lines) and obtained from spline interpolation and differentiation of the
Mz experimental data (dashed lines) shear force and soil reaction distributions at different stages of
loading, for test P32 which was utilized in the calibration

312

1000

Force at Pile Head : kN

Experimental
800

600

400

200

P344 test
0

0.04

0.08

0.12

0.16

0.2

Displacement : m

Force at Pile Head : kN

1000

Calculated

800

600

400

200

P344 test
0

0.04

0.08

0.12

0.16

0.2

Displacement : m

Fig.11: (a) Experimental and (b) computed forcedisplacement curve at sensor DPC for the independentof
calibration test P344 [asymmetric cyclic force, ranging between 960 kN and 0 kN]

313

Bending Moment : kNm


-500
0

1000

2000

3000

Bending Moment : kNm


4000

1000

235 kN

2000

3000

4000

955 kN at

597 kN

12th cycle

Depth : m

-500
0

at 955 kN
6

10

10

12

12

Fig. 12: Comparison of computed (solid lines) and recorded (circles and triangles) bending moment
distributions at different stages of loading, for test P344 [asymmetric cyclic force, ranging between
960 kN and 0 kN]
Soil Reaction : kN / m

Shear Force : kN
-1200
0

-800

-400

400

800

-200
0

597 kN

Depth : m

235 kN

200

400

600

597 kN
at 955 kN

at 955 kN

235 kN
6

10

10

12

12

Fig 13: Comparison of computed (solid lines) and obtained from spline interpolation and differentiation of the
Mz experimental data (dashed lines) shear force and soil reaction distributions at different stages of
loading, for test P344 [asymmetric cyclic force, ranging between 960 kN and 0 kN]

314

1000

Force at Pile Head : kN

Experimental
Calculated

500

-500
P330 test
-1000
-0.2

-0.1

0.1

0.2

Displacement : m
Fig. 14: Comparison of computed (gray solid line) and experimental (black solid line) forcedisplacement
curve at sensor DPC for test P330 [symmetric cyclic force, ranging between 960 kN and 960 kN]

Bending Moment : kNm


-500
0

1000

2000

3000

Bending Moment : kNm


-3000
0

4000

360 kN

602 kN

-1000

500

960 kN
12th cycle

Depth : m

-2000

960 kN
6

2nd cycle

10

10

12

12

Fig. 15: Comparison of computed (solid lines) and recorded (circles and triangles) bending moment
distributions at different stages of loading, for test P330 [symmetric cyclic force, ranging between
960 kN and 960 kN]

315

Soil Reaction : kN / m

Shear Force : kN
-1200
0

-800

-400

400

800

-200
0

602 kN

360 kN

200

400

602 kN

600

at 960 kN

at 960 kN

Depth : m

360 kN
6

10

10

12

12

Fig. 16: Comparison of computed (solid lines) and obtained from spline interpolation and differentiation of the
Mz experimental data (dashed lines) shear force and soil reaction distributions at different stages of
loading, for test P330 [symmetric cyclic force, ranging between 960 kN and 960 kN]

strength relaxation and stiffness hardening of


the pile with cyclic loading, as well as the
reduction in hysteretic damping due to the
development of a relaxation zone around the
upper part of the pile.
The maximum lateral load applied in the
studied tests was in all cases only 1/3 of Pult . A
next phase of this research program will feature
loads up to 80% of Pult . At such intense loading
the nonlinear phenomena described in this
paper will dominate the pile response ; and they
will provide a severe test for the proposed
method.

CONCLUSIONS
A nonlinear Winkler model is applied to
obtain the response of a vertical pile subjected
to cyclic lateral loading during three centrifuge
tests conducted at LCPC, Nantes. A new soil
reactiondisplacement (py) relationship is
utilized to model the soil reaction including
separation and gapping of the pile from the soil.
The model parameters are related to physical
soil properties. Calibrated to match the results
of one centrifuge test, the model is
subsequently utilized to predict the recorded
data of two other tests. A comparison is given
between the calculated py curves with those
proposed by Reese (1974). The model
reproduces adequately the soil yielding
occurring even at very low loading levels. The
bending moment and shear force distribution
with depth are well predicted, whereas there is
a small discrepancy between computed and
measured soil reactions. Finally, the model is
shown to be capable of predicting some
complicated features of the experimental results
arising from counteracting phenomena such as

ACKNOWLEDGMENTS
The analytical and part of the experimental
work of this paper have been conducted with
the financial support of the EU Fifth Framework
Program
:
Environment,
Energy
and
Sustainable Development Research and
Technological Development Activity of Generic
Nature: The fight against Natural and
Technological Hazards, Research Project
QUAKER, Contract number: EVG1CT2002

316

00064. Dr. V. Georgiannou, of the National


Technical University of Athens, performed some
of the laboratory tests on Fontainebleau sand
utilized in our study.

Curras C.J., (2000) : Seismic soilpileinteraction for


bridge
and
viaduct
structures,
Ph.D.
Dissertation, University of California, Davis
Curras C.J., Boulanger R.W. Kutter B.L., and Wilson
D.W. (1999) : Seismic soilpilestructure
interaction in softclay, Proceedings of the 2nd
International Conference on Earthquake
Geotechnical Engineering, P.Seco Pinto (ed.),
published by A.A. Balkema, Rotterdam.

REFERENCES
Angelides D.C & Roesset J.M. (1981) : Nonlinear
lateral dynamic stiffness of piles, Journal of
the Geotechnical Engineering Division, ASCE,
107(11),14431459.

Duncan J. M., Evans L. T., Jr. & Ooi P.S.K. (1994) :


Lateral load analysis of single piles and drilled
shafts,
Journal
of
the
Geotechnical
Engineering Division, ASCE, 120(5), 10181033.

Ashour M. & Norris G. (2000): "Modelling lateral soil


pile response based on soilpile interaction",
Journal of Geotechnical Geoenvironmental
Engineering,ASCE, 126(5), 420-28.

Garnier J. (2002): Properties of soil samples used in


centrifuge models, Invited Keynote Lecture,
International
Conference
on
Physical
Modelling in Geotechnics ICPMG 02, R.
Phillips et al. (Eds), published by A.A. Balkema,
Rotterdam, 1, 5-19.

Baber, T.T. & Wen, Y.-K. (1981): Random vibration


of hysteretic degrading systems, Journal of
Engineering Mechanics, ASCE, 107, 10691087.

Gazetas G. & Dobry R. (1984) : Horizontal


response of piles in layered soils, Journal of
Geotechnical Engineering, ASCE, 110, 2040.

Badoni D. & Makris N. (1995): Nonlinear response


of single piles under lateral inertial and seismic
loads, Soil Dynamics and Earthquake
Engineering, 15, 29-43.

Gerolymos N. & Gazetas G (2005a): Constitutive


model for 1D cyclic soil behavior applied to
seismic analysis of layered deposits , Soils &
Foundations, 45(3), 147-159

Banerjee P.K. (1978, a) : Analysis of axially and


laterally loaded pile groups, Developments in
Soil Mechanics, Applied Science Publishers,
London, 317346.

Gerolymos
N.
&
Gazetas
G
(2005b):
Phenomenological model applied to inelastic
response of soil pile interaction systems,
Soils & Foundations, 45(4), 119-132

Banerjee P.K. & Davies T.G. (1978, b) : The


behaviour of axially and laterally loaded piles
embedded
in
nonhomogeneous
soils,
Geotechnique, 28(3),309326.
Bouc

R. (1971): Modele mathematique


hysteresis,Acustica, 21, 16-25 .

Hansen B.J. (1961) : The ultimate resistance of rigid


piles against transversal forces, The Danish
Geotechnical Institute Bulletin, 12, 59.

Boulanger R.W., Curras C.J., Kutter B.L., Wilson


D.W, and Abhari A (1999) : Seismic
soilpilestructure experiments and analyses,
Journal
of
Geotechnical
and
Geoenvironmental Engineering, ASCE, 125 (9),
750759.
Broms B.B. (1964): Lateral resistance of piles in
cohesionless soils, Journal of Soil Mechanics
and Foundations Division, ASCE, 1964, 90,
SM3, 123-56.

Hardin B. O. (1978): Nature of StressStrain


Behavior for Soils, Proceedings of the
Specialty
Conference
on
Earthquake
Engineering and Soil Dynamics, ASCE, 1, 3
90.
Kim B.T., Kim N.K., Lee W.J., and Kim Y.S. (2004) :
Experimental loadtransfer curves for laterally
loaded piles in NakDong River Sand, Journal
of Geotechnical and Geoenvironmental
Engineering, ASCE, 130(4), 416425.
Kimura M., Adachi T., Kamei H. & Zhang F (1995):
3D finite element analyses of the ultimate
behavior of laterally loaded castinplace
concrete piles, Proc. 15th Int. Symp. on
Numerical Models in Geomechanic., Davos,
Switzerland, 589-94.

Broms B.B. (1964): Lateral resistance of piles in


cohesive soils, Journal of Soil Mechanics and
Foundations Division, ASCE, 1964, 90, SM2,
27-63.
Budhu M. & Davies T.G. (1987) : Nonlinear analysis
of laterally loaded piles in cohesionless soils,
Canadian Geotechnical Journal, 24, 289296.

Kucukarslan S. & Banerjee P.K. (2004) : Inelastic


analysis of pilesoil interaction, Journal of

317

Reese L.C. (1997) : Analysis of laterally loaded


piles in weak rock, Journal of Geotechnical
and Geoenvironmental Engineering, ASCE,
123(11), 10101017

Geotechnical
and
Geoenvironmental
Engineering, ASCE, 130(11), 11521157.
Loh

C.H., Cheng C.R., Wen Y.K. (1995) :


Probabilistic evaluation of liquefaction
potential under earthquake loading, Soil
Dynamics and Earthquake Engineering, 14,
269278.

Reese L.C. (1986) : Behavior of piles and pile


groups under lateral load, Federal Highway
Administration Report
FHWA/RD85/106,
Washington D.C.

Matlock H. (1970): Correlations for Design of


Laterally Loaded Piles in Soft Clay,
Proceedings,
Second
Annual
Offshore
Technology Conference, Houston, Texas, 1,
Paper No. OTC 1204, 577-594.

Reese L. C., Cox, W. R. & Koop, F.D. (1974):


Analysis of laterally loaded piles in sand,
Paper No. OTC 2080, Proceedings, Fifth
Annual Offshore Technology Conference,
Houston, Texas, 1974, II, 473-485.

Matlock H., Foo S.H.C., & Bryant L.M. (1978) :


Simulation of lateral pile behavior under
earthquake motion, Earthquake Engineering
and Soil Dynamics, ASCE, II, 600619.

Reese, L. C., Cox, W. R. & Koop, F. D. (1975):


Field testing and analysis of laterally loaded
piles in stiff clay, Proceedings, Seventh
Offshore Technology Conference, Houston,
Texas, II, Paper No. OTC 2312, 672-690.

McClelland B. & Focht J.A.,Jr, (1958) Soil modulus


for laterallyloaded piles, Transaction of
ASCE, 123, 10491086

Reese L. C. & Van Impe W.F. (2001): Single Piles


and Pile Groups under Lateral Loading, A.A.
Balkema.

Murchison J. M. & O Neil M. W. (1984): Evaluation


of py relationships in cohessionless soil,
Analysis and design of pile foundations, ASCE,
New York, 174-91.

Rosquot F. (2004) : Pieux sous charge latrale


cyclique. Thse de doctorat de l'Ecole Centrale
de Nantes et de l'Universit de Nantes, 305p.

Nogami T., Otani J., Konagai K., & Chen H.I.,


(1992) : Nonlinear soilpile interaction model
for dynamic lateral motion, Journal of
Geotechnical Engineering, ASCE, 118(1),
89106.

Rosquot F., Canepa Y., Garnier J., Thorel L.,


Thtiot N. (2003): Cyclic loading effect on pile
p-y
curves:
centrifuge
modelling,
Foundations : Innovations, observations,
design and practice, ICE, London, Thomas
Telford, 767-775.

O Neil M. W. & Murchison J. M. (1983): An


evaluation of py relationships in sand, Report
of
the
American
Petroleum
Institute,
Washington, D.C.

Rosquot F., Garnier J., Thorel L., Canepa Y.


(2004): Horizontal cyclic loading of piles
installed in sand : Study of the pile head
displacement and maximum bending moment,
Proceedings of the International Conference
on Cyclic Behaviour of Soils and Liquefaction
Phenomena, Bochum, T. Triantafyllidis (Ed.),
Taylor & Francis, 363-368.

Pender M.J. (1993) : A seismic pile foundation


design analysis, Bulletin of the New Zealand
National Society for Earthquake Engineering,
26(1), 49160
Pires J.A., Ang A.HS., Katayama I. (1989) :
Probalistic
analysis
of
liquefaction,
Proceedings of 4th International Conference on
Soil Dynamics and Earthquake Engineering,
Mexico, 1989.

Schofield A.N. (1980): Cambridge geotechnical


centrifuge operations, Geotechnique, 25(4),
743761.
Schofield A.N. (1981): Dynamic and Earthquake
Geotechnical
Centrifuge
Modelling,
Proceedings of the International Conference
on Recent Advances in Geotechnical
Earthquake Engineering and Soil Dynamics, S.
Prakash Foundation, 1081-1100.

Poulos H.G. & Davis EH., (1980) : Pile Foundation


Analysis and Design. John Wiley & Sons.
QUAKER (2002) : Fault-Rupture and Strong Shaking
Effects on the Safety of Composite
Foundations
and
Pipeline
Systems:
http://www.dundee.ac.uk/civileng/quaker

Stevens J.B, & Audibert J.M.E. (1979) :


Reexamination of py curve formulations,
Proceedings of the 11th Offshore Technology
Conference, Paper No OTC 3402, 397403

Randolph M.F. (1981): The response of flexible


piles to lateral loading, Geotechnique, 31(2),
247259.

318

Tabesh A. & Poulos H.G. (2001): The Effects of


Soil Yielding on Seismic Response of Single
Piles, Soils and Foundations, 41(3), 116.
Trochanis A., Bielak J, & Christiano P. (1991):
Threedimensional nonlinear study of piles,
Journal of the Geotechnical Engineering,
ASCE, 117, GT3, 429-47.
Trochanis A., Bielak J., & Christiano P. (1994):
Simplified model for analysis of one or two
piles, Journal of Engineering Mechanics,
ASCE, 120, 308-29.
Wakai A., Gose Sh., & Ugai K. (1999): 3D
ElastoPlastic Finite Element Analyses of Pile
Foundations subjected to Lateral Loading,
Soils & Foundations, 39(1), 97111.
Wen Y.K. (1976): Method for random vibration of
hysteretic systems, Journal of Engineering
Mechanics, ASCE, 102, 249-263.
Whitman R.V., Lambe P.C., Kutter B.L. (1981) :
Initial Results from a Stacked Ring Apparatus
for Simulation of a Soil Profile. International
Conference
on
Recent
Advances
in
Geotechnical Earthquake Engineering and Soil
Dynamics, S. Prakash Foundation, III,
11051113.
Wu B., Broms B., & Choa V., (1998): Design of
Laterally Loaded Piles in Cohesive Soils Using
py curves, Soils & Foundations, 38(2), 1726.
Yegian M., & Wright S.G. (1973) : Lateral soil
resistancedisplacement relationships for pile
foundations in soft clays, Proceedings of the
5th Offshore Technology Conference, Paper No.
OTC 1893, II663II676.

319

Seismic Retrofitting of the Piled Foundation of a


Reinforced Concrete Building
M. Maugeri, F. Castelli
Department of Civil and Geo-environmental Engineering
University of Catania, Italy

Abstract
The Sicilian earthquake of 13th December, 1990 (Italy) was not very strong, but it caused
considerable damages on structures. Particularly at the Augusta site located on the east
coast of Sicily, because the soft soil foundation, soil amplification was observed and
reinforced concrete buildings were severely damaged. To evaluate the possibility to repair
these buildings, an investigation on soil and structures was carried out, including loading
tests on piles of the piled foundation. In the paper the results of these experimental
investigations are reported and analyzed with the aim to derive important considerations
regarding the behaviour of the piles subject to additional seismic loads. Finally the
retrofitting of the piled foundation is described.

INTRODUCTION

Seismology deals with the estimation of the


seismic action at the conventional bedrock;
Geotechnical Engineering deals with the
analysis of the soil stability and the site local
effects for the definition of an elastic response
spectrum at the soil surface; Structural
Engineering allows us to evaluate an anelastic
design spectrum, which is obtained by the
elastic response one taking into account the
structural ductility.
For the common structures the seismic
action is given by the National Regulations. For
important structures the deterministic evaluation
of the seismic action could be more
appropriated.
The deterministic evaluation of the seismic
action can be achieved by means of the use of
recorded accelerograms conveniently scaled in
amplitude and frequency in order to simulate
the scenario earthquake, or, for strategic,
monumental and historical buildings, by means
of synthetic accelerograms developed from the
modelling of the source mechanisms, related to
the scenario earthquake, of medium or high
magnitude.
For common structures the local seismic
response can be evaluated according to
Regulations and Recommendations, such as

The seismic risk mitigation is one of the


greatest challenges of the Civil Engineering in
the Third Millennium. An important contribution
toward this challenge can be given by the
Geotechnical Engineering.
As stressed by Hudson (1981), the soilstructure interaction is a crucial point for the
evaluation of the structure seismic response, so
Earthquake Engineering must be founded on
Geotechnical Engineering.
In the Geotechnical Earthquake Engineering
the researches are essentially devoted to two
different fields.
The first field, which is the most traditional,
deals with the modelling of the behaviour of the
single geotechnical construction and/or soilstructure interaction.
The second field deals with the stability of
physical environment. This last field involves,
more than the first, different scientific sectors,
because
the
Geotechnical
Earthquake
Engineering must interact with the Seismology,
the Geology and the Structural Engineering.
The common goal is the evaluation of the
seismic action that shall hit civil engineering
structures.

320

Eurocode 8 (EC8-Part 1, 2003). Nevertheless,


the soil factor due to possible amplification
phenomena, given by Regulations and/or
Recommendations, on the basis of medium
values of soil properties (Vs,30), can be widely
exceeded for particular soil profile conditions.
Thus, for important structures, it is appropriated
to evaluate the seismic local ground response,
by means of 1-D, 2-D and/or 3-D codes, taking
into account the soil profile, up to a significant
depth, and the non linear soil behaviour.
An accurate geotechnical characterization is
fundamental for evaluating the seismic local site
response, as well as for modelling the soil
sample behaviour and for analyzing the stability
of the site where the building foundations rest
on. The geotechnical investigations must be
planned in relation to the specific design.
The geotechnical characterization must
include the estimation of the soil dynamic
properties, nevertheless it must be based on an
accurate estimation of the soil static properties.
For common structures the dynamic
geotechnical characterization can be in part
performed through empirical correlations, which
join the dynamic soil properties to the static
ones. These correlations can be utilized by the
designers, if they are validated by means of the
results related to some test sites.
For
important
structures,
strategic,
monumental or historical, it is necessary to
perform specific dynamic in-situ and laboratory
tests, in relation to the specific geotechnical
problem to be solved.
When the geotechnical characterization
phase is completed, before to analyze the
foundation behaviour, it is necessary to
investigate on the overall stability of the site
where the foundations rest on, as stressed by
Eurocode 8 (EC8-Part 5, 2003). In this context,
the most important phenomena to be analyzed
are the slope stability and the liquefaction.
Evaluated the stability of the site where the
buildings rest on, it is possible to investigate the
behaviour of shallow and deep foundations in
relation to the ultimate limit state and to the
damage limitation state in static and/or dynamic
conditions.
For foundations built in seismic areas, the
demands to sustain load and deformation
during an earthquake will probably be the most
severe in their design life.
Today the seismic evaluation and design of

soil-foundation interaction is an area of active


research and seismic analysis should be
performed to evaluate loading conditions during
and after a seismic event.
With the aim to analyze and to empathize
the importance of the different phases of a
seismic design process, in the paper the
approach adopted for the seismic retrofitting of
a group of reinforced concrete buildings (Fig 1)
located in Augusta (Italy), damaged by the
Sicilian earthquake of 13th December 1990, is
described and discussed.

Fig 1: View of building damaged due to earthquake

The site was well investigated by means of


laboratory and in-situ tests including seismic
tests (Fig 2).
Seismic action has been evaluated by the
deterministic approach (Frenna & Maugeri,
1995; Maugeri & Frenna, 1995) based on
accelerograms registered during the 13th
December 1990 earthquake, scaled to simulate
the effects of the 1693 scenario earthquake.
By the dynamic response analysis of soil a
site depending response spectrum has been
determined. Then a design spectrum more
severe than that adopted by the Italian Seismic
Regulation has been derived.

Fig 2: Layout of the site with the location of buildings


and in situ tests

321

For evaluating the possibility to repair the


buildings, an investigation on soil, structures
and foundations was carried out.
As buildings are founded on piles (Fig 3),
vertical and horizontal loading tests on two
single pile tests were conducted. Also tests on
a three-piles group were carried out.

derived from the laboratory tests.


Cross-hole and down-hole tests indicate an
average value of the shear waves velocity
equal to Vs = 320 m/s up to -10 m from the
ground surface. An average value of the shear
waves velocity approximately equal to Vs = 530
m/s can be estimated up to -30 m from the
ground surface.
The small strain shear modulus of soil Go
can be considered an essential parameter for
estimation of the soil response, especially for
dynamic loading conditions.
The modulus Go was determined both by insitu than by laboratory tests. In laboratory the
shear modulus and the damping ratio (D) was
determined by Resonant Column tests (RCT)
and cyclic loading torsional shear tests
(CLTST), even if it was attempted to assess Go
also by means of empirical correlations based
on in-situ test results.
By laboratory tests, the values of Go were
determined at shear strain levels less than
0.001 %. It is possible to observe that quite a
good agreement exists between the laboratory
and in-situ test results.
In particular, the values of Go for the upper
silty-clay steadily increase from 20 to 80 MPa
with depth. In the transition zone, where stiff
sand layers are present, Go increases up to 110
MPa (Cavallaro et al., 1996). In the lower bluegrey clay Go values range in the range 80 up to
120 MPa.
A small difference between the values of
shear modulus Go determined by this last
procedure and the values derived by resonant
column tests can be observed in the first 10 m
below the ground surface (Fig 5).

Fig 3: Detail of the pile foundation

Loading tests showed that the seismic


actions have not damaged the effectiveness of
the soil-pile system, as the bearing capacity of
the piles appeared unchanged in spite of the
additional loads applied on the structures during
the earthquake.
Nevertheless, because the existing piled
foundations were enable to carry on the
additional seismic actions, a number of piles
were added.
The procedure discussed and analysed
permit to define a state of practice in the field of
seismic retrofitting and/or strengthening of pile
foundations.
SOIL GEOTECHNICAL CHARACTERIZATION

TESTS ON PILES

The investigated area has plane dimensions


of 4100 m2 (Fig 2).
A detailed description of the soil properties
is reported by Castelli & Maugeri (2004).
The thickness of the deposits varies from 50
up to 300 m. The upper part of these deposits
mainly consist of alternating layers of grey-siltyclay and sandy-clay which is locally called
Augusta Clay. Layers of sand were found at
depths of between 9 and 12 m.
The lower deposits mainly consist of a
medium stiff, over-consolidated marine clay with
low to medium plasticity index.
Fig 4 shows the main index parameters and
mechanical properties of the Augusta clay

The site under consideration, the Saline of


Augusta, is one of the most seismically active
areas of Italy. The city of Augusta has been
destroyed by three disastrous earthquakes with
an MKS intensity ranging from IX up to XI
(Postpischl, 1985).
Even if 1990 earthquake was of moderate
magnitude (M = 5,4 and MKS intensity equal to
VII), it caused 19 victims and severe damages
to buildings (Fig 6) and infrastructures.
The reinforced concrete buildings under
consideration are founded on bored piles
having a length of 20 meters and a diameter of
500 mm.

322

- 0.80 m

Sand backfill
- 3.0 m

Silty - clay
Clayey - silt

- 9.0 m

Sandy - clay
- 12.0 m

Blue - grey
clay

- 60.0 m

Fig 4: Soil properties derived by laboratory tests

a good agreement with the value of 1,53 MN


computed in undrained conditions by the wellknown Vesics (1977) relationships.
To evaluate the pile performance due to
lateral loads, after conducting the axial loading
tests, two lateral loading tests on single pile
were also performed (Figs 11 and 12).
The ultimate failure lateral load derived by
the load-deflection curves results greater than
20 kN.

To check the integrity of the piled


foundations after the earthquake, the piles were
inspected (Fig 7) and an experimental loading
tests program was carried out.
Two single piles of the piled foundation were
chosen and their head was disconnected from
the cap to perform vertical (Fig 8) and horizontal
loading tests (Fig 9). At the end of the loading
tests, the connection between the piles of the
foundation and the cap was restored (Fig 10).
A detailed description of the results of the
loading tests carried out on piles is reported by
Castelli & Maugeri (2004).
Synthetically, according to the procedure
suggested by Chin (1970), the vertical ultimate
failure load derived by the loading tests is
ranging between around 1,5 MN and 1,7 MN, in
G0 [MPa]
0

50

100

150

0
10

Fig 6: Damages observed on concrete structures

Profondit [m]

20
30
40
50
60
70
80

CH Test
RCT
RCT-Ismes
MLTST
CLTST

90

Fig 5: Values of the shear modulus Go derived by in


situ and laboratory tests

Fig 7: Piles inspection

323

used with little computational effort.


These methods have given results that are
often in remarkable agreement with the
mathematical models (Novak, 1974; Markis &
Gazetas, 1992; Abghari & Chai, 1995; Tabesh
& Poulos, 2001).
Pseudo-static approaches, as example, for
the seismic analysis and design of foundations
are attractive for design engineers because
they are simple, when compared to difficult and
more complex dynamic analyses.
There are two sources of loading of the pile:
inertial loading of the pile head, caused by the
lateral forces imposed on the structure by the
earthquake and which are then imposed on the
piles, and kinematic loading along the length

Fig 8: View of a loading test on pile

Horizontal Load [KN]


0.0

10

12

14

16

18

0.2

Lateral Deflection

[mm]

0.4

Fig 9: Horizontal loading test on pile

0.6
0.8
1.0
1.2
1.4
1.6

Test no.1

1.8
2.0

Fig 11: Single pile horizontal loading tests results


Horizontal Load [KN]
0.0

12

15

18

21

0.5

[mm]

1.0

Lateral Deflection

Fig 10: View of the pile-cap reconnection

The loading tests carried out on a threepiles group (Fig 13) subjected to horizontal load
(Fig 14) furnished the results reported in Fig 15
in terms of load-deflection curve.

1.5
2.0
2.5
3.0
3.5
4.0

PROPOSED PROCEDURE

4.5

Test no.2

5.0

In the last years several simplified


approaches for the analysis of single piles or
pile groups have been developed that can be

Fig 12: Single pile horizontal loading tests results

324

Horizontal Load [kN]

of the pile, imposed by the lateral ground


movements developed by the earthquake.
Generally in the pile design only the effects
of inertial loading are considered, even if
kinematic loading is also very important
(Gazetas, 1984; Tazoh et al., 1988; Masayuki &
Shoichi, 1991; Kavvadas & Gazetas, 1993).
Part 5 of the Eurocode 8 states that piles
shall be designed for the following two loading
conditions:

25

50

75

100

125

150

175

200

Lateral Deflection [mm]

1
2
3
4
5
6

Pile Group

Fig 15: Piles group horizontal loading tests results

- a) inertia forces on the superstructure


transmitted on the heads of the piles in
the form of axial and horizontal forces
and moment;
- b) soil deformations arising from the
passage of seismic waves which impose
curvatures and thereby lateral strain on
the piles along their whole length.
(a)

While there is ample experience of carrying


out the equivalent static analysis for the inertial
loading (type (a)), no specific method or
procedures
are
available
to
predict
deformations and bending moment from the
kinematic loading (type (b)).
To take into account these two sources of
loading, Abghari & Chai (1995) proposed an
analysis in which the pile was subjected to the
free-field soil displacements at each node along
its length (Fig 16).
These displacements were obtained from a
separate free-field site response analysis. The
inertial forces acting on the pile were obtained
from the product of mass and spectral
acceleration. These forces were applied to the
pile as static forces.
Similarly Tabesh & Poulos (2001) proposed
a simple approximate methodology for
estimating the maximum internal forces of piles
subjected to lateral seismic excitation.
The method involves the evaluation of the
maximum free-field soil movements caused by
earthquake computed, as example, by the well
known SHAKE (Schnabel et al., 1972) program
or similar computer code and the analysis of the
response of the pile to the maximum free-field
soil static movements plus a static loading at
the pile head (Poulos,2006), which depends on
the computed maximum spectral acceleration of

(b)

Fig 13: Plan (a) and three-dimensional (b) view of


the pile group

Fig 14: Horizontal loading tests on piles group

325

the structure being supported.


According to this approach in the paper a
pseudo-static push over analysis using a p-y
method is adopted to simulate the behaviour of
a single pile and/or an individual pile of a group
subjected to lateral soil movements (kinematic
loading) and static loadings at the pile head
(inertial loading).
The response of the complete system
constituted by soil, foundation and over
structure is obtained by the overlapping of the
effects, even if, the validity of the principle of
the overlapping of the effects is correct only in
the hypothesis of linear behaviour of all
components (Kausel & Roesset, 1974; Gazetas
& Milonakis, 1998).
Nevertheless, the overlapping of the effects
can be considered acceptable for systems
moderately non linear (Milonakis et al., 1997).

represented by a series of nonlinear p-y curves


that vary with depth and soil type.
The p-y method is still widespread in
practice and based on literature review it
appears to be one of the most attractive
methods for civil engineers.
Researches based on results of field tests
on full scale piles, both in cohesive than in
cohesionless soil, suggests to employ non
linear p-y relationship (Reese et al., 1974;
Reese & Welch, 1975; Reese & Van Impe,
2001; Juirnarongrit & Ashford, 2006), thus in
the proposed approach a hyperbolic p-y
relationship has been adopted:

p( z ) =

yp( z )
y (z)
1
+ p
Esi ( z ) plim ( z )

(1)

where Esi [FL-2] is the initial modulus of


horizontal sub-grade reaction, yp [L] is the pile
lateral deflection, p and plim [FL-2] are the
mobilized and the ultimate horizontal soil
resistance respectively.
The hyperbolic p-y relationship (1) is defined
by the two parameters plim and Esi. By this
approach a situation of layered soil can be
easily analyzed assuming for these parameters
different values along the pile length.
The numerical model is based on an
iterative procedure taking into account the
decrease of the model stiffness with the
increasing of the applied horizontal load
(Castelli et al., 1995; Castelli & Maugeri, 1999;
Castelli 2002; 2006).
The response of a laterally loaded pile group
with relatively closely spaced piles is different
from that of a single pile, because of the
interaction
between
piles
through
the
surrounding soil.
Experiments have been carried out in the
last decade with the aim of deriving general
rules to adapt p-y curves to take into account
the group effects.
Brown et al., (1988), as example, introduced
the term shadowing to mean the phenomenon
for which the soil resistance of a pile in a trailing
row is reduced because of the presence of the
leading pile ahead of it.
Brown et al., (1988) defined the concept of
p-multiplier fm, a multiplier of the limiting values
plim capable of stretching the p-y curve for the
single pile to account for the interaction among

Free - Field
Soil Movement

Pile
Deflection

Pile
Deflection

Fig. 16: Profile of external soil movement caused by


earthquake
INERTIAL LOADING

Several approaches are currently available


to analyze the behaviour of piles subjected to
lateral load ranging from complex models, as
non linear dynamic analysis by 2D or 3D finite
element methods, to the use of simplified
approaches as limit equilibrium (LE) approach
and p-y analysis approach.
The p-y approach for analyzing the
response of laterally loaded piles is essentially
a modification of the basic Winkler model,
where p is the soil pressure per unit length of
pile and y is the pile deflection. The soil is

326

used to adjust the magnitude of load carried by


each row of piles in the group.
According to Fig 17 it should be expected
that the resulting initial modulus of horizontal
sub-grade reaction of a pile group is softer than
the value of an isolated pile (Ashour et al.,
2004).
To adapt the hyperbolic p-y relationships
defined for a single pile to the case of a pile
group, a multiplier should be considered also
for the initial modulus of horizontal sub-grade
reaction Esi.
In practice, the use of this multiplier is
similar to the traditional approach given by
NAVFAC (1982) in which the modulus of subgrade reaction is reduced by a factor taken as a
function of pile spacing (a value equal to 1 is
suggested at 8 diameter pile spacing, varying
linearly to 0,25 at 3 diameters).
The modulus of horizontal sub-grade
reaction reflects the soil-pile interaction at any
level of pile loading or soil strain.
Then, Esi will account for the additional
strains in the adjacent soil due to pile
interaction within the group and, consequently,
Esi (i.e. the secant slope of the p-y curve
reported in Fig 17) of an individual pile in a
group will be reduced (Ashour et al., 2004).
Consequently we can assume for a pile
group:

the piles in a group.


This effect is related to the influence of the
leading row of piles on the yield zones
developed in the soil ahead of the trailing row of
piles. Because of this overlapping of failure
zones, the front row will be pushing into virgin
soil while the trailing row will be pushing into
soil which is in the shadow of the front row
piles.
A consequence of this loss of soil resistance
for piles in a trailing row is that the leading piles
in a group will carry a higher proportion of the
overall applied load than the trailing piles. This
effect also results in gap formation behind the
closely spaced piles and an increase in group
deflection.
It has been shown both theoretically and
experimentally that the shadowing effect
becomes less significant as the spacing
between piles increases and is relatively
unimportant for centre-to-centre spacing greater
than about six pile diameters (Cox et al., 1984;
Brown & Shie, 1991; Ng et al., 2001).
Similar to the design of axially loaded piles
using a efficiency method, in a pile group
subjected to horizontal load, the p-y
relationships for the individual piles are modified
to take into account the group effects by
stretching the curve in the direction of
deflection.
To extend the method of the p-y curves to
the case of a pile group, the p-multiplier
concept proposed by Brown et al., (1988) has
been applied. The multiplier has obviously
values in the range 0 to 1.
As suggested by Brown et al., (1988), a
given row, within the group, could be
represented by a single pile p-y curve multiplied
by a load factor to account for row position.
With this aim, in the proposed approach the
p-y curve of the piles belonging to each row of
the group is obtained by scaling down the p-y
curve of the single pile, multiplying the values of
plim by a group reduction factor named fm.
The generally accepted approach is to
assume that p-multipliers are constant with
depth. That is, a constant p-multiplier is applied
to the set of p-y curves for all depths in a given
pile row. Thus, individual p-y curves for a pile
are adjusted by the same amount, regardless of
variations in the soil profile or the depth below
the ground surface.
Experimentally determined multipliers are

p( z ) =

yp(z)
yp(z)
1
+
m E si ( z ) f m p lim ( z )

(2)

where m is the empirical factor introduced in


the proposed method to extend the single pile
analysis to the case of a pile group (Castelli,
2006).
Also the multiplier m has obviously values
in the range 0 to 1 and it can be defined as the
ratio between the initial modulus of horizontal
sub-grade reaction of the pile in a group
(Esi)group and that of the single pile (Esi)single :

(Esi)group

m = ___________

(3)

(Esi)single
A rather large amount of experiments have
been carried out in the last decade for the
determination of the p-y multipliers.

327

Lateral resistance p

p lim (z)

interfaces to a corresponding set of discrete


relative motions between piles and free-field
soil.
The implementation of the method involves
imposing a known free-field soil movement
profile.
When the expected free-field movement is
large enough to cause the ultimate pressure of
laterally spreading soils to be fully mobilized,
the ultimate pressure, instead of free-field soil
movement, may be used.
Piles surrounded by moving soil are relevant
for many engineering applications.
This soil movement will in turn displace the
pile a certain amount depending on the relative
stiffnesses between the pile and the soil.
The soil loading must be considered by
taking into account the relative movement
between the soil and the pile.
If the soil mass moves and the pile movement
yp is less than the soil movement ys, the soil
exerts a driving force on the pile.
However, if the pile movement yp is greater
than the soil movement ys, the soil provides the
resistance force plim to the pile.
The response of the pile can then
obtained by solving the following governing
differential equation:

single pile

E si(z)

p(z)
pile group

p(z) = fm p(z)single
E si(z) = m E si (z)single

Lateral deflection y

Fig 17: fm and m multipliers for group effects

The value of the group reduction factors


seems do not be depend by the soil type, pile
type and load level, thus it is reasonable to
suggest values only in terms of number of piles,
pile spacing and position in the group (McVay
et al., 1998; Rollins et al., 1998; Zhang et al.,
1999; Rollins et al., 2006).
The p-y multipliers can be assumed
constant for each row and independent of the
number of piles contained in the rows, so in the
practice a unique multiplier for each row is
generally fixed.
The empirical factor m can be backcalculated comparing the numerical results with
experimental evidences (Castelli, 2006).
The advantage of the proposed approach is
that the load-deflection response of the pile
group can be calculated using solutions for the
response of a single pile.
Naturally, this approach is based on an
estimation of only the average head lateral
deflection of the pile group.

EI

d4y
p( y p y s )
dz 4

(4)

where EI is the pile stiffness, p is the soil


reaction per unit length of pile and z is the
depth.
This equation can be solved by a numerical
procedure based on a pile finite element
discretization, in which the pile load (force per
unit area [FL-2]) due to relative pile-soil
movement (yp ys), can be represented by a
series of p-y curves on both sides of the pile
shaft for all its length.
For a simple approach an idealized elastic
p-y relationship could be used:

KINEMATIC LOADING

The passage of seismic waves through the


soil surrounding a pile imposes lateral
movements and curvatures on the pile,
generating kinematic bending moments even in
absence of a over-structure.
The simplified numerical approach adopted
in the paper is based on the idea to couple and
to evaluate the effects due to the applied load
at the pile head and lateral movements along
the pile length by a series of independent p-y
curves, relating soil reaction and relative soil pile movements.
Since these relations are non linear, the
equivalent linear procedure using secant moduli
is normally used to establish a law relating a set
of discrete interaction forces at the soil/pile

p( z ) =

E si ( z )
[ y p ( z ) y s ( z )]
D

(5)

where D is the pile diameter.


To take into account that the lateral pile
response to static or dynamic loading is non
linear, the following p-y relationship could be
adopted:

328

p( z ) =

[ y p ( z ) y s ( z )]
| yp(z) ys(z) |
1
+
E si ( z )
p lim ( z )

area, to analyse the possibility that even more


intense earthquake than the one recorded on
13th December, 1990 could occur, the
accelerations and the frequencies of the
recordings at Sortino E-W have been scaled.
The recording was scaled using the
maximum
acceleration
0,3g
and
the
predominant period of earthquake T = 0,5 sec.
The model for the evaluation of the seismic
soil response (Maugeri & Frenna, 1987)
requires the knowledge of the depth of the
conventional bedrock. In this case it has been
placed 80 m below the ground surface.
The numerical analysis provides the timehistory response of displacements, velocity and
accelerations. Fig 19 shows the profiles of the
numerical results versus depth (Maugeri &
Frenna, 1995; Frenna & Maugeri, 1995).

(6)

in which Esi, yp, p and plim have the same


meaning of equation (1).
According to the proposed approach, when
an individual pile of a group is taken into
consideration, in eq (6) the empirical reduction
factors fm and m should be used:

p( z ) =

[ y p ( z ) y s ( z )]
| y p(z) ys(z) |
1
+
m E si ( z )
f m p lim ( z )

(7)

The approach proposed was employed and


implemented in an original computer code.
Comparison with field measurements, shows
that this procedure can give lateral deflection,
bending moment and shear force distributions
on piles which agree with experimental ones
(Castelli & Maugeri, 2007).
SEISMIC SOIL RESPONSE

Augusta site, situated in the South-Eastern


part of Sicily (Italy), is placed on one of the
most intense seismic hazard areas of Italy.
The estimation of damages caused by the
13th December 1990 South-Eastern Sicily
earthquake, has emphasized the effects of the
phenomenon of local amplification, which
caused serious damages to concrete type of
structures.
The response of the soil, schematized as a
one-dimensional system with several degrees
of freedom and lumped masses, has been
evaluated with a hysteretical-simplified model,
through the integration of the system of
differential equations of the dynamic equilibrium
with variable coefficients, using the Newmarks
numerical integration method step-by-step
(Maugeri & Frenna, 1987). The model is
capable to take into account the decay of the
elastic shear modulus with strain.
The excitation at the base of the model were
the accelerograms of the 13th December, 1990
earthquake recorded around Sortino E-W,
which is the most representative among those
made on rock (Fig 18).
As it has been emphasized in paragraph 1,
related to the seismic characteristics of the

Fig 18: Accelerogram recorded at Sortino E-W


COMPARISON WITH MEASURED BEHAVIOUR
OF PILES UNDER LATERAL LOADING

The model has been implemented in an


original computer code (Castelli & Maugeri,
2007).

Fig 19: Results of the seismic soil response

329

assumed equal to 37 MN/m2.


The single piles were in free-head condition
while the pile group was in fixed-head condition.
As the horizontal load during the loading test
was applied at a distance from the ground
surface of 43 cm, the numerical analysis has
been carried out considering the pile loaded
with a horizontal force and a moment at the
head.
In Fig 20 is reported the comparison
between measured (Test no.1) and computed
single pile head lateral deflection and a good
agreement can be observed between
experimental and numerical results.
To reduce the computed load-carrying
capacity of the single pile to the pile group
rows, according to the design curve showing pmultiplier values reported by Rollins et al.,
(1998), a p-multiplier fm equal to 0,4 was
assumed.
The values of the initial modulus of
horizontal sub-grade reaction of the piles group
were reduced assuming for the empirical factor
m the same value of fm.
Also in this case a good agreement between
numerical results and experimental evidences
can be observed (Fig 21).
Loading tests showed that the seismic
actions have not damaged the effectiveness of
the soil-pile system, as the bearing capacity of
the piles appeared unchanged in spite of the
additional loads applied on the structures during
the earthquake.

The code allows the assessment of the


lateral response (deflection, moment and shear
force distribution) of an isolated pile and a pile
group including the p-y curve along the length
of the isolated pile or the individual piles in the
group.
The isolated pile or the individual pile of a
group can be considered subjected to the
simultaneous application of a lateral force and
moment at its head and a free-field soil
movement profile along its length.
Validation of the proposed approach has
been carried out comparing the numerical
results with those measured in loading tests,
both on single pile and piles group. Model
parameters have been assessed to obtain
numerical results close to those measured.
The numerical analysis was carried out for
the single piles and then for the pile groups,
considering in this last case the experimental
curve in terms of average lateral load per pile
versus average pile head lateral deflection.
A difficulty of applying the proposed p-y
curve approach is the appropriate evaluation of
the functions parameters for a realistic
estimation of single pile and/or pile group
performance.
The adoption of hyperbolic p-y curves, in
particular, requires the determination of the
ultimate horizontal soil resistance plim and initial
modulus of horizontal sub-grade reaction Esi.
In a single pile-soil interaction, the ultimate
horizontal soil resistance plim can be evaluated
according to the well known formulas existing in
literature (Matlock, 1970) both for cohesive
(Broms, 1964a) than for cohesionless soils
(Broms, 1964b).
As concern the initial modulus of horizontal
sub-grade reaction Esi, the relationships
proposed by Matlock (1970), Welch & Reese
(1972), Robertson et al., (1989) for hyperbolic
p-y curves can be used.
Taking into account the profile of the values
of the undrained shear strength derived from in
situ and laboratory tests, the numerical analysis
has been carried out with values of plim ranging
between 0,024 MN/m2 from the ground surface
up to 1,33 MN/m2 at the pile tip.
The initial modulus was: Esi = (Esoi + ki z),
being z the depth, ki equal to 19,75 N/cm3 the
gradient of the initial modulus of horizontal subgrade reaction and Esoi the initial modulus at the
ground surface, that in the analysis was

Orizzontal Load [kN]


0.0

10

12

14

16

Single Pile
Lateral Deflection [mm]

0.3

0.6

Computed
Measured
0.9

1.2

Lenght: L = 20 m
Diameter: D = 0.50 m
Young Modulus:
E = 20 x 103 MPa

1.5

Fig 20: Comparison between experimental and


computed load-deflection curve for the
single pile

330

Horizontal Load [kN]


0

10

20

30

40

50

60

effects are considered, the maximum moments


along the pile length can be underestimated.
When combined effects are taken into
consideration, an increasing of the maximum
bending moment can be observed, with a
considerable reduction of the safety margin
respect to the estimated plasticity moment of
the pile equal to My = 120 kNm.
As the existing piled foundation was enable
to carry on the additional seismic actions
(kinematic and inertial), a number of piles were
added (Fig 23).

70

Pile Group
Lateral Deflection [mm]

Computed
Measured

Lenght: L = 20 m
Diameter: D = 0.50 m
Reduction Factors:
fm and m = 0.40

CONCLUDING REMARKS

The paper describe the results of an


experimental investigation carried out on a
group of reinforced concrete buildings located
on the East coast of Sicily (Italy) damaged by
the Sicilian earthquake of 13th December, 1990.
To evaluate the possibility to repair these
buildings, an investigation on soil, structures
and piled foundations was carried out.
To test the integrity of the piles after the
earthquake, loading tests were carried on. In
particular, two single piles and a three-piles
group beneath the existing foundation were

10

Fig 21: Comparison between experimental and


computed load-deflection curve for the
three-piles group
EVALUATION OF INERTIAL AND
KINEMATIC EFFECTS

To evaluate the influence of the inertial and


kinematic effects, the numerical analysis has
been carried out via the pseudo-static approach
following the proposed procedure (Castelli &
Maugeri, 2007).
To simulate the kinematic effects, the
displacement profile of the free-field soil
movements (Fig 19) evaluated by the seismic
soil response analysis (Maugeri & Frenna,
1995) has been considered.
So that numerical simulation provides
bending moment distribution along the pile
length. Two cases have been considered:

Pile Moment (KNm)


0

-5

10 15 20 25 30 35 40 45 50

2
Combined Effects (H = 15 kN)

Combined Effects (H = 12.2 kN)

Depth (m)

- lateral inertial load but kinematic ground


movements not included in the analysis
(inertial effects only);
- lateral inertial load ranging from 12,2 kN to
15 kN (combined effects).

Inertial Effects (H = 15 kN)


Inertial Effects (H = 12.2 kN)

yp

ys

10

12

Pile
Movement

H
i
Soil
Pressure

14

The numerical analyses relative to this last


case of combined effects (kinematic and
inertial) have been performed according to p-y
relationship defined by equation (7).
The results obtained for the single pile are
summarized in Fig 22.
Analyzing these results it is possible to
observe that, as expected, if kinematic effects
(lateral movements) are ignored and only
inertial (lateral load applied at the pile head)

Free - Field
Soil Movement
"p - y" spring

16
18
20
-5

10 15 20 25 30 35 40 45 50

Fig 22: Numerical results for combined effects on


single pile (kinematic and inertial)

331

Broms, B.B. (1964b) Lateral resistance of piles in


cohesionless soils, Journal of Soil Mechanics
and Foundation Engineering Div., ASCE, Vol.XCI,
No.SM-3, pp.123-156

chosen and their head was disconnected from


the cap to execute horizontal loading tests.
At the end, the connection between the piles
and the cap was restored.
For the seismic retrofitting of the structures,
the internal response of the piles subjected to
earthquake loading, has been studied.
The numerical simulation has been carried
out via a pseudo-static approach which involves
the evaluation of the free-field soil movements
caused by earthquake.
As the existing piled foundation was enable
to carry on the additional seismic actions
(kinematic and inertial), a number of piles were
added to the piled foundation.

Brown, D.A., Morrison, C., and Reese, L. C. (1998)


Lateral load behaviour of pile group in sand,
Journal of Geotechnical and Geo-environmental
Engineering, ASCE, Vol.114, No.11, pp.12611276
Brown, D.A., and Shie, C.F. (1991) Modification of
P-Y curves to account for group effects on
laterally loaded piles, ASCE Geotechnical
Congress 1991, Reston, Va., Vol.1, pp.479-490
Castelli, F. (2002) Discussion on Response of
laterally loaded large-diameter bored pile groups
by C.W.W. Ng, L. Zhang & D.C.N. Nip, Journal of
Geotechnical and Geo-environmental Eng.,
ASCE, Vol.128, No.11, pp.963-965

ACKNOWLEDGEMENTS

Funding for this research was provided by


DPC-Reluis Research Project no.6.4 Deep
Foundations.

Castelli, F. (2006) Non linear evaluation of pile


groups lateral deflection. Proc. International
Conference on Piling and Deep Foundations,
Amsterdam, 31 May - 2 June, 2006, pp.127-134

REFERENCES

Castelli, F., Maugeri, M. and Motta, E. (1995)


Modelling of a horizontal loading test on a pile
beneath a building damaged due to earthquake,
Proc. II National Conference of Earthquake
Engineering, Siena, Italy, Vol.I, pp.185-194 (in
Italian)

Abghari, A., and Chai, J. (1995) Modelling of soilpile superstructure interaction for bridge
foundations, Performance of deep foundations
under seismic loading, J. P. Turner ed., ASCE,
New York, pp.4559

Castelli, F., Maugeri, M. and Motta, E. (1995) Non


linear analysis of the deflection of a pile
subjected to horizontal loads, Italian Geotech.
Journal, Vol.XXIX, No.4, pp.289-303 (in Italian)

Ashour M., Pilling, P., and Norris, G. (2004) Lateral


behavior of pile groups in layered soils, Journal
of Geotechnical and Geo-environmental Eng.,
ASCE, Vol.130, No.6, pp.580-592

Castelli, F., and Maugeri, M. (1999) Nonlinear soil


effects in the design of pile group in pseudostatic conditions, Proc. IX National Conference
Ingegneria Sismica in Italia, Torino, Italy, p.10
(in Italian)

Broms, B.B. (1964a) Lateral resistance of piles in


cohesive soils, Journal of Soil Mechanics and
Foundation Engineering Div., ASCE, Vol. XC,
No.SM-2, pp.27-63

Fig 23: Pile foundation with new piles added

332

Castelli, F., and Maugeri, M. (2004) Analysis of the


behaviour of the piled foundations of a group of
earthquake damaged buildings, Proceedings V
International Conference on Case Histories in
Geotechnical Engineering, New York, 13-17 April,
2004, paper No.1.45, p.6
Castelli, F., and Maugeri, M. (2007) A pseudo-static
analysis for the evaluation of the lateral
behaviour of pile groups, Proc. 4th International
Conference on Earthquake Geotechnical Eng.,
Thessaloniki, Greece, Paper No.1399

Kavvadas, M., and Gazetas, G. (1993) Kinematic


seismic response and bending of free-head piles
in layered soil, Gotechnique, Vol.43, No.2,
pp.207-222
Hudson, D.E. (1981) The role of Geotechnical
Engineering in earthquake problems, Proc.
International Conference on Recent Advances in
Geotechnical Earthquake Engineering and Soil
Dynamics. St. Louis
Mc Vay, M., Zhang, L., Molnit, T., and Lai, P. (1998)
Centrifughe testing of large laterally loaded pile
groups in sands, Journal of Geotechnical and
Geo-environmental Engineering, ASCE, Vol.124,
No.10, pp.1016-1026

Cavallaro,
A.
and
Maugeri,
M.
(1996)
Comportamento tensionale e deformativo
dellargilla di Augusta sottoposta a carichi ciclici,
Ingegneria Sismica, Vol.XIII, pp.30-40 (in Italian)

Masayuki, H., and Shoichi, N. (1991) A study on


pile forces of a pile group in layered soil under
nd
seismic loading, Proceedings 2 International
Conference
on
Recent
Advances
in
Geotechnical Earthquake Engineering and Soil
Dynamics, St. Louis

Chin, F.K. (1970). Estimation of ultimate load of


piles not carried to failure, Proc. 2nd SE Asian
Conference on Soil Eng., Singapore, pp.81-92
Cox, W. R., Dixon, D. A., and Murphy, B. S. (1984)
Lateral load tests on 25.4-mm (1-in.) diameter
piles in very soft clay in side-by-side and in-line
groups. Laterally loaded deep foundations:
analysis and performance, ASTM STP 835, J. A.
Langer, E. T. Mosley, and C.D. Thompson (ed),
American Society for Testing and Materials, pp.
122-139

Matlock, H. (1970) Correlations for design of


laterally loaded piles in soft clay, Proceedings II
Offshore Technical Conference, Houston, Texas,
Vol.1, pp.577-594
Maugeri, M., and Frenna, S.M. (1987) Modello
isteretico semplificato per la determinazione
della risposta dei terreni in campo non lineare,
Proceedings III National Conference Ingegneria
Sismica in Italia, Vol.II, pp.169-288 (in Italian)

Eurocode 8 - Part 5 (2003) Design of structures for


earthquake resistance Part 5: Foundations,
retaining structures and geotechnical aspects,
European Prestandard, ENV1998, European
Comm. for Standardization, Brussels, Belgium

Maugeri, M., and Frenna, S.M. (1995) Soilresponse analyses for the 1990 South-East Sicily
rd
earthquake, Proc. 3 International Conference
on Recent Advances in Geotechnical Earthquake
Eng. and Soil Dynamics, St. Louis, Vol.II,No.9.17,
pp.653-658

Frenna, S.M., and Maugeri, M. (1995) Microzoning


of the Saline site of Augusta town (Sicily, Italy),
th
Proc. 10 International Conference on Earth.
Engineering, Vienna, Vol.I, pp.43-49
Gazetas, G. (1984) Seismic response of endbearing single piles. Soil Dynamics Earthquake
Engineering, Vol.3, No.2, pp.82-93
Gazetas, G., and Mylonakis, G. (1998) Seismic soilstructure interaction: new evidence and
emerging issue, Geotechnical Earthquake
Engineering and Soil Dynamics III, ASCE, eds. P.
Dakoulas, M. K. Yegian, and R. D. Holtz, Vol.II,
pp.1119-1174

Markis, N., and Gazetas, G. (1992) Dynamic pilesoil-pile interaction. Part II: Lateral and seismic
response,
Earthquake
Engineering
and
Structural Dynamics, Vol.21, pp.145-162
Mylonakis, G., Nikolaou, A., and Gazetas, G. (1997)
Soil-pile-bridge seismic interaction: kinematic
and inertial effects. Part I: soft soil, Earthquake
Engineering Structural Dynamics, Vol.27, No.3,
pp.337-359

Juirnarongrit, T., and Ashford, S.A. (2006) Soil-Pile


response to blast-induced lateral spreading. II:
Analysis and Assessment of the p-y method,
Journal of Geotechnical and Geo-environmental
Engineering, ASCE, Vol.132, No.2, pp.163-172

Ng, C.W.W., Zhang, L. and Nip, D.C.N. (2001)


"Response of laterally loaded large-diameter
bored pile groups, Journal of Geotechnical and
Geo-environmental Engineering, ASCE, Vol.127,
No.8, pp.658-669

Kausel, E., and Roesset, J.M. (1974) Soil Structure


Interaction for Nuclear Containment Structures,
Proceedings ASCE, Power Division Specialty
Conference, Boulder, Colorado

Novak, M. (1974) Dynamic stiffness and damping of


piles, Canadian Geotechnical Journal, Ottawa,
Vol.11, pp.574-598

333

Postpischl, D. (1985) Atlas of isosismal maps of


Italian Earthquake, Italian National Research
Council (CNR), Geodynamical Project, Roma

Vesic, A. S. (1977) Design of pile foundations,


Transport Res. Board., Report No.42, p.68
Welch, R.C., and Reese, L.C. (1972) Laterally
loaded behavior of drilled shafts, Research
Report No.3-5-65-89, Center for Highway
Research, University of Texas

Poulos, H.G. (2006) Ground movements A hidden


source of loading on deep foundations, John
Mitchell Lecture, Proceedings International
Conference on Piling and Deep Foundations,
Amsterdam, pp.2- 19

Zhang, L., Mc Vay, M., and Lai, P. (1999) Numerical


analysis of laterally loaded 3x3 to 7x3 pile groups
in sands, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol.125,
No.11, pp.936-946

Reese, L.C., Cox, W.R., and Koop, F.D. (1974)


Analysis of laterally loaded piles in sand, Proc.
VI Offshore Technology Conference, Houston,
Texas, pp.473-483
Reese, L.C., and Welch, R.C. (1975) Lateral
loading of deep foundations in stiff clay,
Geotechnical Testing Journal, GTJODJ, Vol.XII,
No.1, pp.30-38
Reese, L.C., and Van Impe, W.F. (2001) Single
piles and pile groups under lateral loading,
Balkema, Rotterdam, The Netherlands
Robertson, P.K., Davies, M.P., and Campanella,
R.G. (1989) Design of laterally loaded driven
piles using the flat dilatometer, Geotechnical
Testing Journal, GTJODJ, Vol.12, No.1, pp.3038
Rollins, K.M., Peterson, K.T., and Weaver, T.J.
(1998) Lateral load behavior of full-scale pile
group in clay, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol.124, No.6,
pp.468-478
Rollins, K.M., Olsen, K.G., Jensen, D.H., Garrett,
B.H., Olsen, R.J., and Egbert, J.J. (2006) Pile
Spacing Effects on Lateral Pile Group Behavior:
Analysis, Journal of Geotechnical and Geoenvironmental Engineering, Vol.132, No.10, pp.
1272-1283
Schnalbel, B., Lysmer, J., and Seed, H. B. (1972)
SHAKE, a computer program for earthquake
response analysis of horizontally layered sites,
Rep. No.EERC 72-12, Earthquake Engineering
Res., Ctr., National Information System for
Earthquake Engineering
Tabesh, A. and Poulos, H.G. (2001) Pseudo-static
approach for seismic analysis of single piles,
Journal of Geotechnical and Geo-environmental
Engineering, ASCE, Vol.127, No.9, pp.757-765
Tazoh, T., Wakahara, T., Shimizu, K., and Matsuzaki,
M. (1988) Effective motion of group pile
th
foundations, Proceedings 9 World Conference
Earthquake Engineering, Tokyo, Vol.3, pp.587592
US Navy, (1982) Foundations and earth retaining
structures design manual, NAVFAC,Department
of Navy, DM 7.2, Alexandria, Washington D.C.

334

Influence of Batter Piles on the Seismic Response of Pile Groups


A. Giannakou, N. Gerolymos , G. Gazetas
National Technical University of Athens, Greece

Abstract
The behavior of inclined piles embedded in a soil layer with a linearly increasing stiffness
under seismic loading is investigated through the use of 3D finite element analyses. The
structure is modeled as a singledegreeoffreedom system composed of a concentrated
mass and a column. Three simple groups of two piles are studied: (a) one comprising a
vertical pile and a pile inclined at 25o, (b) one consisting of two piles symmetrically inclined
at 25o and (c) a group of two vertical piles. Two piletocap connections, fixed and hinged,
are analysed. The influence of the key parameters is analysed and non-dimensional
diagrams are presented to illustrate the role of raked piles on pile group response.

INTRODUCTION

accumulating
revealing
the
successful
performance of inclined piles. Characteristic
examples include the Maya Warf (Kobe
earthquake 1995), the Ohba Ohashi bridge in
Japan (1984) and the Landing Road bridge
(New Zealand) in the Edgecumbe earthquake.
Sadek & Shahrour (2006) studied the seismic
response of inclined micropiles, and showed
that, for kinematic loading, a group of four
symmetrically inclined micropiles exhibits lower
values of lateral acceleration at the cap level
and larger values of internal forces in the piles
compared to a group of vertical micropiles.
In this paper 3D finite element analyses are
performed to study the response of three simple
pile group configurations which contain batter
piles, embedded in a non-homogenous soil: (a)
a group comprising of a vertical and a 25o
inclined pile, (b) one of two piles inclined at 25o,
and (c) one of two vertical piles. The latter is
used as a reference for comparison with the
batter pile groups for detecting the beneficial or
detrimental role of pile inclination to the seismic
behaviour of the foundation. Analyses are
performed with and without the presence of a
superstructure so that both the kinematic and
the total response of the piles is investigated.
Three real acceleration time histories are used
as base excitation. Since the type of pile-to-cap
connection has shown to greatly influence the
performance of the foundation, analyses were
conducted for both hinged and fixed conditions.

Little attention has been paid to the


influence of inclined piles in the seismic
response of pile groups. Such piles find
frequent use in foundations when substantial
lateral stiffness is required. For years, the
seismic behavior of inclined piles has been
considered detrimental, and many codes
require that such piles be avoided. For
instance, the French Seismic Code (AFPS 90)
states flatly that inclined piles should not be
used to resist seismic loads. The seismic
Eurocode EC8 / Part 5, dealing with
geotechnics and foundations, is a little less
restrictive, stating: It is recommended that no
inclined piles be used for transmitting lateral
loads to the soil. If, in any case, such piles are
used, they must be designed to carry safely
axial as well as bending loading.
In recent years evidence has been
accumulating that inclined piles may, at least in
certain cases and if properly designed, be
beneficial rather than detrimental both for the
structure they support and the piles themselves
(Gazetas and Mylonakis, 1998). Recent
research (Guin, 1997) has shown that the
response of a typical bridge type structure to
representative seismic excitation may improve
in many respects when supported by inclined
piles. The Edgecumbe 1987 case history offers
some support, although indirect, to such
findings. In addition, field evidence has been

335

PROBLEM DEFINITION AND MODEL


DESCRIPTION

elements. Due to symmetry, only half the model


is analysed, thus significantly reducing
computation time. Fig 2 depicts the finite
element discretization for the three pile group
configurations studied. Two piletocap fixity
conditions are considered, namely fixed-head
and hinged-head piles. Rayleigh damping is
used to model material damping, taken equal to
5% for the range of frequencies between the
eigenfrequency of the soil deposit and the
dominant frequency of the earthquake.
Appropriate kinematic constraints are imposed
to the lateral edges of the model allowing the
latter to move as a laminar box.
Three real acceleration time histories
covering a wide range of frequencies were used
as base excitations (Fig 3): (i) Lefkada record,
PGA = 0.42g, from the Lefkada 2003
earthquake (M = 6.4), (ii) Aegionbedrock, PGA
= 0.39 g from the Aegion 1995 earthquake (M =
6.2), and (iii) JMA record, PGA = 0.83 g, from
the Kobe 1995 earthquake (M = 7.2).

Fig 1 shows the characteristics


of
the
analysed pile group configurations. All piles of
Youngs modulus Ep, diameter d, length L, and
cross-sectional moment of inertia Ip, are
considered to be linear elastic.
m = 200 ton

L = 15 m

hst = 12 m

25o

s(z)

d=1m

KINEMATIC RESPONSE OF GROUPS WITH


INCLINED PILES

Distributions of displacements and internal


forces for the Lefkada (2003) record are only
presented in this paper due to space limitations
(Fig 4), but the observed trends and
conclusions are valid for all earthquake records
used in this study. Moreover, the maximum
kinematic response of the configurations to all
three acceleration time histories utilised as base
excitation is presented in Fig 5. All results are
normalized by the total (kinematic and inertial)
response of the fixed-head vertical pile group,
since this is the type of pile foundation that is
commonly used in practice.
The great advantage of the group
configurations containing batter piles is that the
maximum lateral displacement at the pile-cap is
smaller than that of the vertical pile group.
Particularly in the case of the group with two
inclined piles the displacements are up to 40%
less than in the case of the group of vertical
fixed-head piles. This observation matches well
the frequently observed behavior of the inclined
piles in the field under conditions of extensive
soil movements (such as liquefaction and lateral
spreading). For the group of hinged-head
inclined piles the displacement at the pile-cap is
only 20% less than that of the group with fixedhead vertical piles.

Fig 1: Problem geometry and soil profile


studied
The center-to-center distance between the
piles at the pile head level, s, is 3 m. The piles
are embedded in a soil deposit with
fundamental period Ts = 0.29 sec, whose
Youngs modulus, Es, is assumed to vary
linearly with depth according to the following
equation:
z
Es = Eo
(1)
d
where Eo is the Youngs modulus at depth z = d.
It is noted that non-homogenous profiles with Es
linearly increasing with depth are a reasonable
assumption for normally consolidated clays
(Velez et al, 1983).
The problem is analyzed in 3-D, making use
of the Finite Element (FE) method. Both the pile
and the soil are assumed to be linear elastic
materials. The superstructure is modelled as a
singledegreeoffreedom
oscillator
with
characteristics of a typical bridge, i.e. a
concentrated mass mst = 200 ton (deck), and a
column with height hst = 12 m (pier). Its fixed
base fundamental period is Tst = 0.45 s. Soil
and pile are modelled with eight-noded brick

336

Configuration a

Configuration b

Configuration c

Fig 2: Finite element discretization of the three pile group configurations analyzed
0.5

0.8
0.6

0.4

JMA (1995)

0.4

a:g

a:g

0
5

10

15

20

25

30

t : sec

-0.4

0
-0.1 0

15

t : sec

-0.4

0.82 g

-1

-0.5
3

0.5

0.39 g

0.4
0.3

Aegion Bedrock
(1995)
SA : g

Lefkada

0.1
1

JMA (1995)

2.5

0.2

a:g

10

-0.3

-0.8

-0.2

0.1

-0.2

-0.6

0
-0.1 0

Lefkada (2003)

0.2

0.2
-0.2 0

0.42 g

0.3

1.5

Aegion Bedrock
(1995)

1
0.5

t : sec

Aigio
JMA
(2003)
Lefkada

-0.3

0.5

1.5

T : sec

Fig 3: Acceleration response spectra

order of 6 pile diameters), and is smaller than


the maximum bending moment of the vertical
pile group.
Inclined piles in both configurations b and
c develop significantly larger (in the order of 4)
axial force than that of the vertical group for
both fixed and hinged pile-head conditions. This
force, however, occurs at a depth of
approximately 10 pile diameters.

The asymmetric configuration b exhibits


systematically greater bending moments at the
connection with the cap than the piles of the
other two groups, mainly due to the asymmetry
of the foundation. As a result, the inclined pile,
which is stiffer than the vertical one, resists
more the soil movement and, hence, greater
bending moments are developed, as depicted
in Fig 5 for all records utilized herein. The
inclined pile in the symmetric group
configuration exhibits similar behavior with the
vertical pile in the asymmetric pile group
configuration, and both develop their maximum
bending moment two diameters below the
ground surface. When the piles are hinged to
the cap, the bending moments at the pile-head
become zero, while the inclined piles in
configuration c develop their maximum
bending moment at a greater depth (in the

SOILFOUNDATIONSTRUCTURE INTERACTION

Results obtained with the superstructure are


presented in Figs 6-9. The floor acceleration
response spectra at the mass of the
superstructure normalized by the peak ground
acceleration of the free field are depicted in Fig
6 for fixed and hinged pile-to-cap connection.

337

Hinged pile-to-cap connection

Fixed pile-to-cap connection

Uk (z) / max Uk+I, fixed-head vertical group


0

0.4

0.8

1.2

0.4

0.8

1.2

(a)

z/d

4
6
8

Vertical group

10

Asymmetric group
Vertical pile

12

Asymmetric group
Inclined pile

14

Symmetric group

16

Mk (z) / max Mk+i, fixed-head vertical group


-0.2
0

0.2

0.4

0.6

0.8

-0.2

0.2

0.4

0.6

0.8

(b)

z/d

4
6
8
10
12
14
16

Nk (z) / max Nk+i, fixed-head vertical group


00

0.1

0.2

0.3

0.1

0.2

0.3

22
44

(c)
z/d

66
88
100
122
144
166

Fig 4: Normalized distributions of kinematic : (a) horizontal displacement, (b) bending moment, and (c) axial
force along the pile at the time when the maximum occurs. Normalization with respect to their peak
maximum total counterparts of the group of rigidly connected vertical piles.(Excitation: Lefkada
record)

338

fixed-head vertical group

max Uk / max Uk+i,

1.5
fixed pile-to-cap
connection
1.0

0.5

Configuration :

max Mk / max Mk+i,

fixed-head vertical group

Vertical

b
Vertical

(a)

fixed pile-to-cap
connection

0.8

Inclined

hinged pile-to-cap
connection

0.6

0.4

0.2
0
a

b
Vertical

fixed-head vertical group

b
Inclined

1.0

Configuration :

max Nk / max Nk+i,

hinged pile-to-cap
connection

Inclined

b
Vertical

(b)

Inclined

0.4
fixed pile-to-cap
connection

hinged pile-to-cap
connection

0.3

0.2

0.1

Configuration :

b
Vertical

Inclined

b
Vertical

Inclined

(c)

Fig 5: Normalized maximum kinematic response: (a) peak maximum horizontal displacement, (b) peak
maximum bending moment, and (c) peak maximum axial force along the pile. Normalization with
respect to their peak maximum total counterparts of the group of rigidly connected vertical piles
(maximum, minimum, and average values from the three accelerograms)

339

inclined piles is significantly smaller than that of


the vertical group for both types of pile-to-cap
connection analysed. This reduction in the
structural distress is attributed to the smaller
pile-cap horizontal displacements and to the
larger pile-cap rotations of the groups with
inclined piles.

The hinged connection acts as a seismic


isolation not only for the foundation but also for
the superstructure, resulting in a significant
decrease of spectral acceleration values at the
deck, and smaller distress of the piles and the
pier compared to that of the group with vertical
piles (Fig 7). The horizontal drift when the
superstructure is supported by groups with

Hinged pile-to-cap connection

Vertical group
Asymmetric group
Symmetric group

SA

mass

/ PGA free-field

Fixed pile-to-cap connection

0
0

0.5

1.5

0.5

1.5
T : sec

max uh stru / max uh stru, vertical group

Fig 6: Normalized floor response spectra of the three group configurations with respect to the free-field peak
ground acceleration (Excitation: Lefkada record)
1.6
fixed pile-to-cap
connection

hinged pile-to-cap
connection

1.2

0.8

0.4

Configuration :

Fig 7: Normalized maximum drift of the superstructure with respect to its maximum counterpart of the group of
rigidly connected vertical piles (maximum, minimum, and average values from the three
accelerograms).

role of inclined piles in the reduction of


displacements attenuates significantly.
The bending moment that develops at the
cap when the group contains inclined piles is
significantly larger than in the vertical group and
can be up to 2.5 times greater in the case of
configuration c for fixed connection (Figs 8b
and 9b).

Configurations with inclined piles exhibit smaller


displacements than the vertical group, as
depicted in Figs 8a and 9a. In the case of the
group with two inclined piles the displacements
are up to 40% lower. For the asymmetric
configuration this percentage reduces 30%.
However, changing the degree of fixity from
fixed- to hinged-head condition the beneficial

340

Hinged pile-to-cap connection

Fixed pile-to-cap connection

Uk+i (z) / max Uk+i, fixed-head vertical group


0

0.2

0.4

0.6

0.8

1.2

0.2

0.4

0.6

0.8

1.2

2.5

1.2

(a)

z/d

4
6
8

Vertical group

10

Asymmetric group
Vertical pile

12

Asymmetric group
Inclined pile

14

Symmetric group

16

Mk+i (z) / max Mk+i, fixed-head vertical group


-0.5
0

0.5

1.5

2.5

-0.5

0.5

1.5

(b)

z/d

4
6
8
10
12
14
16

Nk+i (z) / max Nk+i, fixed-head vertical group


0

0.2

0.4

0.6

0.8

1.2

0.2

0.4

0.6

0.8

2
4

(c)
z/d

6
8
10
12
14
16

Fig 8: Normalized distributions of total (kinematic + inertial) : (a) horizontal displacement, (b) bending moment,
and (c) axial force along the pile at the time when the maximum occurs. Normalization with respect to
their peak maximum total counterparts of the group of rigidly connected vertical piles.(Excitation:
Lefkada record)

341

1.5
fixed pile-to-cap
connection
1.0

0.5
hinged pile-to-cap
connection
0
Configuration :

max Mk+i / max Mk+i,

fixed-head vertical group

Vertical

b
Vertical

(a)

Inclined

hinged pile-to-cap
connection

2.5
2
1.5
1
0.5

fixed pile-to-cap
connection

0
a

b
Vertical

fixed-head vertical group

Configuration :

max Nk+i / max Nk+i,

b
Inclined

Inclined

b
Vertical

(b)

Inclined

1.5
fixed pile-to-cap
connection

hinged pile-to-cap
connection

1.0

0.5

Configuration :

b
Vertical

Inclined

b
Vertical

Inclined

(c)

Fig 9: Normalized maximum total (kinematic + inertial) response: (a) peak maximum horizontal displacement,
(b) peak maximum bending moment, and (c) peak maximum axial force along the pile. Normalization
with respect to their peak maximum total counterparts of the group of rigidly connected vertical piles
(maximum, minimum, and average values from the three accelerograms)

342

Fifth Framework Program : Environment,


Energy
and
Sustainable
Development
Research and Technological Development
Activity of Generic Nature: The fight against
Natural and Technological Hazards, Research
Project QUAKER, Contract number: EVG1CT
200200064.

The hinged connection leads to a moderate


decrease in bending moment for the case of the
vertical group, but significant for the case of the
configurations containing batter piles for both
profiles.
As presented in Fig 9c, the developing axial
forces in configurations with inclined piles are
smaller than in the vertical group for all profiles
studied (40% for configuration c and 30% for
configuration b for fixed connection). The fixity
condition at the cap level does not seem to
significantly influence the developing axial
force, except in the case of configuration c,
where the axial force is significantly reduced at
the pile-cap level (Fig 8c).

REFERENCES
AFPS (1990) Association franaise de Gnie
Parasismique. Recommandations AFPS 90.
Presses des Ponts et Chausses.
Eurocode EC8 (1994) Structures in seismic regions,
Part 5: Foundations, retaining structures, and
geotechnical aspects.

CONCLUSIONS

Gazetas G. and Mylonakis G. (1998) "Seismic soilstructure interaction: new evidence and
emerging issues, " Geotechnical Earthquake
Engineering and Soil Dynamics III, ASCE, 2,
Geotechnical Special Publication, 75, pp 11191174.

The seismic response of pile group


configurations containing batter piles is
analysed in this paper. Parametric numerical
analyses are performed to gain insight on the
response of inclined piles to seismic loading.
The response of pile groups to three different
acceleration time histories, with two different
pile-to-cap connections is studied. The results
of the numerical analyses showed that groups
with batter piles provide large lateral resistance
which results in smaller lateral displacements
during earthquake loading.
Generally, the symmetric group with inclined
piles in conjunction with hinged pile-to-cap
connection results in the most satisfactory
performance of both the superstructure and the
foundation. The hinged connection acts as a
seismic isolation not only for the foundation but
also for the superstructure, resulting in smaller
distress to both the piles and the pier compared
to that of the group with vertical piles. However,
attention should be drawn to the fact that
groups with inclined piles tend to develop large
bending moments at the connection with the
pile-cap. Thus, detailing of the pile-cap
connection must be sufficient to undertake
safely this bending moment and to assure
adequate inelastic deformation during large
earthquakes. This is not an easy task however,
given the difficulty in achieving ductile
behaviour of a connection with batter pile.

Guin J. (1997) "Advances in soil-pile-structure


interaction and non-linear pile behavior," PhD
thesis, State University of New York at Buffalo.
Sadek M. & Shahrour I. (2006) "Influence of the
head and tip connection on the seismic
performance of micropiles," Soil Dynamics and
Earthquake Engineering, Vol. 26, pp 461-468.
Velez A., Gazetas G., and Krishnan R. (1983)
"Lateral dynamic response of constrained-head
piles," Journal of Geotechnical Engineering,
ASCE, Vol. 109.

ACKNOWLEDGMENTS
The work presented in this paper has been
conducted with the financial support of the EU

343

Effect of the Natural Frequencies of the Coupled Soil-Pile-Structure


System on Pile Dynamic Response
E. Rovithis1, E. Kirtas1, K. Pitilakis1
1

Department of Civil Engineering, Aristotle University of Thessaloniki, Greece

Abstract
The important role of soil-pile-structure interaction on the seismic response of pilesupported structures has been widely established by several theoretical and experimental
studies. However, due to the complex nature of the physical phenomenon, the identification
of the key parameters that affect the interaction mechanism remains a crucial task. Along
these lines, the frequency content of the seismic motion constitutes an important parameter
that affects the intensity of soil-pile-structure interaction phenomena and consequently the
seismic response of pile-supported structures. In this paper, soil-pile-structure interaction is
numerically examined in the frequency domain in order to obtain an insight on the effect of
the natural frequencies of the system on pile and structural dynamic response. Kinematic
soil-pile interaction is initially investigated under harmonic excitation applied at the base of
the soil profile, emphasizing on the effect of the soil fundamental frequencies on the pile
bending moments. Within the framework of the validation of the adopted soil-pile model,
additional analyses were performed implementing pilehead loading in order to obtain pile
impedance functions and compare them to existing analytical formulas. In a second stage of
analysis the coupled soil-pile-structure system is examined, adopting single pile-supported
structures with different dynamic characteristics. Soil-pile-structure interaction effects are
discussed in terms of the fundamental frequency of the superstructure, the bending
moments generated along the pile shaft and the pile head motion. The analysis results
indicate that for certain soil-pile cases, the kinematic peak pile bending moments may occur
at frequencies higher than the fundamental frequency of the soil deposit. On the other hand,
the investigation of the coupled soil-pile-superstructure system revealed that structural
oscillations transmit large bending moments on the pilehead when the modified frequency of
the structure due to SSI effects is close to the fundamental frequency of the input motion.
Furthermore, the combined effect of inertial and kinematic interaction may under certain
conditions result in a significant amplification of the pilehead horizontal motion at the
fundamental frequency of the pile-structure system.

INTRODUCTION

simplified interaction models such as the Beam


on Dynamic Winkler Foundation approach as
well as more rigorous BEM or FEM
formulations. The majority of these methods
utilize
the
principle
of
superposition,
decomposing therefore the total interaction
mechanism into kinematic and inertial effects
(Kagawa and Kraft, 1982, Nogami, 1983,
Gazetas, 1984, Fan et al. 1991, Gazetas et al,
1992, Kaynia and Novak, 1992, Makris and
Gazetas, 1992, Gazetas & Mylonakis, 1998).
Although
the
implementation
of
the
superposition approach is by definition
restricted to linear response, it may be a
reasonable approximation even when moderate

During the last two decades significant


progress has taken place in the field of soil-pilestructure
interaction.
Well-documented
centrifuge and shaking table tests (Wilson,
1998, Meymand, 1998) as well as postearthquake investigations on pile failures
(Mizuno, 1987, Tokimatsu et al. 1998, Berill and
Yasuda, 2002) have elucidated some of the
dominant parameters that affect pile and
structural response during seismic shaking.
Moreover, several numerical and analytical
methods have been proposed for the analysis
of soil-pile-structure systems, implementing

344

stage of analysis, the dynamic response of the


coupled soil-pile-structure system is examined,
with emphasis on the role of the dynamic
characteristics of the superstructure. Results
are discussed in terms of the fundamental
frequency of the superstructure, the bending
moments generated along the pile shaft and the
pile head motion.

non-linear soil behavior is anticipated


(Mylonakis et al., 1997). An alternative
approximation to this multi-step method is the
direct analysis of the coupled soil-pile-structure
system (Lok et al. 1998, Guin and Banerjee,
1998), which allows one to capture
simultaneously the interaction effects between
all the components of the system and
additionally permits a more precise evaluation
of the non-linear soil behavior.
The
aforementioned
theoretical
and
experimental studies have established the
significant influence of the frequency content of
the input motion on the seismic response of
pile-supported structures. Both kinematic and
inertial interaction effects present a strong
frequency dependency, which accordingly
affects the structural dynamic characteristics as
well as the incident wave field imposed on the
structure. Therefore, the identification of the
role of the natural frequencies of the system on
the soil-pile-structure interaction is required
towards the prediction of the seismic response
of pile-supported structures, especially in
correlation with the frequency content of the
input motion. Along these lines, resonance
phenomena that may be generated during
seismic shaking could modify substantially the
interaction mechanism, imposing large seismic
loads on the pile-structure system.
In this paper, soil-pile-structure interaction is
numerically examined in the frequency domain,
emphasizing on the effect of the soil and the
structure fundamental frequencies on the
dynamic response of the system. The adopted
analysis procedure comprises two stages. In
the first stage, pile dynamic response due to
kinematic soil-pile interaction is investigated,
utilizing a single pile embedded in a
homogeneous soil stratum of finite depth. The
effect of the soil fundamental frequencies on
the distribution of the steady state bending
moments along the pile shaft is investigated
and discussed emphasizing on the particular
case where the input motion excites higher soil
modes. Kinematic interaction factors are also
computed for both fixed and free pilehead.
Within the framework of the validation of the
adopted soil-pile model, the soil-pile system
was reanalyzed implementing harmonic loading
applied on the pilehead, in order to obtain pile
impedance functions and compare them to
existing analytical expressions. In the second

KINEMATIC RESPONSE ANALYSIS OF THE


SOIL-PILE SYSTEM

Based on the superposition approach, the


analysis of the complete soil-pile-structure
system under seismic excitation is performed
with the following three consecutive steps
(Makris et al. 1994, Gazetas & Mylonakis,
1998): (a) Estimation of the response at the
pilehead (the so-called "foundation input
motion") in the absence of the superstructure's
mass which includes translational and rotational
components (b) Calculation of the dynamic
impedances (springs and dashpots) associated
with each vibration mode of the foundation
(swaying, rocking and cross swaying rocking
oscillation) and (c) Computation of the
superstructure's seismic response supported on
the springs and dashpots of step b and
subjected to the foundation input motion
obtained from step a. Kinematic soil-pile
interaction that is examined in step (a) holds
therefore a significant role in the determination
of the seismic response of pile supported
structures since it can modify substantially the
pilehead motion with respect to the free field
response. Furthermore, soil deformations
during seismic wave propagation may impose
significant kinematic induced bending moments
on the pile foundations (Tazoh et al. 1987).
In order to examine kinematic interaction
effects, an end bearing pile of circular cross
section embedded in a 30m thick homogeneous
soil layer over rigid bedrock was employed
herein. The shear wave velocity of the soil was
considered constant with depth and equal to
Vs=200m/sec that corresponds to soil type C
according to EC8 (CEN 2002). The Poisson
ratio and the hysteretic damping ratio of the soil
were taken equal to v=0.33 and s=0.05
respectively. The relative soil-pile stiffness was
considered equal to Ep/Es=1000 while the pile
slenderness, i.e the ratio of the pile length L to
the pile diameter D, was selected equal to
L/D=20. The dynamic response of the soil-pile

345

system was investigated in the frequency


domain. The analyses were performed with the
general purposed FE code ANSYS (Ansys,
2000), implementing a 3D finite element model
of the soil-pile system. The soil stratum was
meshed with 8-node solid elements while the
pile was modeled with linear elastic beam
elements attached to the FE mesh, thus
ignoring in a first stage any potential soil-pile
gapping mechanism. Introducing elementary
boundaries at the lateral sides of the model
revealed minimal interference to the free field
soil response during the numerical analysis
while the FE mesh size was adequately defined
based on the anticipated wavelength of the
shear waves propagating in the soil.

Kinematic interaction factors

Due to kinematic soil-pile interaction the


horizontal displacement at the pile head up is
generally different from the free field motion uff,
while in the case of free head conditions the
pile head is subjected to an additional rotational
motion p. The kinematic interaction effects, in
terms of the pilehead response, are usually
described by the kinematic interaction factors
(Gazetas, 1984):

Iu =

u ff

, I

p r

(2)

u ff

where r=D/2 is the pile radius. The pilehead to


free field response ratios defined in Equation 2
were numerically obtained for both free and
fixed pile head conditions and were then
compared to kinematic interaction factors
proposed by Fan and Gazetas, 1991 and
Kaynia and Novak, 1992 for a similar soil-pile
system. The comparative results shown in Fig 2
confirm further the accuracy of the obtained
analyses,
highlighting
certain
kinematic
interaction effects with respect to the excitation
frequency as well as to the pilehead conditions.
It is observed that in the case of the fixed head
pile, the kinematic soil-pile interaction has a
dominant effect mainly in the high frequency
range, where the inability of the relatively stiff
pile of the present study to follow the free field
deformation caused a significant reduction of
the pilehead displacement with respect to the
soil motion. On the other hand, when the
rotational degree of freedom is released at the
pilehead adopting free pilehead conditions, the
Iu values exceed unity in the low to intermediate
frequency range that implies a pilehead motion
greater than the free field response. In the high

Soil response analysis

In order to prove that wave propagation


effects are adequately captured by the
employed FE model, soil response analyses
were initially performed under harmonic
displacement applied at the base of the soil
profile. Based on the one-dimensional wave
propagation theory (Roesset, 1977), for a
monochromatic SV wave motion ug(t)=ugeit
introduced at the bedrock level and a
homogeneous soil stratum, the steady state
displacement at the ground surface equals to
uff(t)=uffeit. The amplification function, AF,
defined as the ratio of the amplitude of the
motion at the free surface uff to the amplitude at
the bedrock level ug can be computed from the
equation:
AFsoil () =

up

2
ff =
(1)
ug
Vs*
Vs*
exp(i
) + exp(-i
)
H
H

where H is the thickness of the soil layer and


is the complex shear wave velocity of the form
Vs* = Vs 1 + 2 i s .The amplification function
obtained from the harmonic response analysis
of the soil profile was compared to the
theoretical solution given by Equation 1. Fig 1
shows the aforementioned comparison where a
very good agreement is observed, thus
verifying that soil response is sufficiently
reproduced by the FE model. The specific
frequencies where the soil response is amplified
correspond to the fundamental frequencies of
the soil deposit, which in the particular case are
equal to 1.6Hz, 5Hz and 8.33Hz respectively.

Amplitude

14
12

Numerical AF()

10

Theoritical F()

8
6
4
2
0
0

4 5 6
f(Hz)

9 10

Fig 1: Amplification function of the examined soil


deposit

346

0.5

0.4
0.2

ao

0
0

0.1

0.2

0.3

0.4

FE analysis
Fan and Gazetas
Kaynia and Novak

1.5
1
0.5

ao

0.5

FE analysis
Fan and Gazetas

0.6

1.5

0.8

FE analysis
Fan and Gazetas

0.1

0.2

0.3

0.4

ao

0
0.5

0.1

0.2

0.3

0.4

0.5

(a)
(b)
(c)
Fig 2: Kinematic interaction factors for: a -b) Free head pile c) Fixed head pile
(Homogeneous soil stratum, Ep/Es=1000, L/D=20, s=0.05, =0.33)

the role of the fundamental frequencies of the


soil deposit on the pile response. The
distribution of the steady state bending moment
amplitude along the pile length was therefore
computed at each one of the three fundamental
frequencies of the considered homogeneous
soil layer. The results obtained for free as well
as for fixed pilehead conditions are shown in
Fig 3. It is noted that the distributions of the
bending moments presented in Fig 3 are
normalized to the maximum bending moment
developed along the pile length. More
specifically the bending moment profiles
obtained for the free head pile are normalized
to the maximum moment that occurred at the
3rd natural frequency of the soil deposit (the
exact location is indicated in Fig 3 with a blue
circle) while the bending moments computed for
the fixed head case are normalized to the

frequency range though, the observed


reduction of the pilehead motion with respect to
the free field displacement is less severe than
in the fixed head case. Furthermore, the
rotation of the pilehead is increasing with
frequency. Similar findings regarding the effect
of the pilehead conditions on the kinematic
interaction factors are also reported in the study
of Gazetas et al., (1992).
Kinematic Pile Bending moments

Although kinematic induced pile bending


moments are usually significant in the presence
of sharp stiffness discontinuities in the soil
profile (Mylonakis, 2001), large soil curvatures
may occur also in the case of a homogeneous
soil stratum due to the "wavy" shape that
characterizes the higher response modes of the
soil (Kavvadas and Gazetas, 1993). The latter
is further investigated herein, emphasizing on

M/Mm ax
0.5

M/Mm ax
0.5

10

10

15

Free Head

15

10

Location of
Mmax for
fixed head pile

15

Fixed Head
20

20

(a)

z/ D

z/ D

z/ D

M/Mm ax
0.5

Location of
Mmax for
free head pile

20

(b)

(c)

Fig 3: Effect of pile head conditions on kinematic bending moments amplitude obtained at the a)
First b) Second and c) Third natural frequency of the soil deposit

347

maximum moment obtained at the 2nd natural


frequency of the soil (red circle in Fig 3).
Reviewing the results, it is observed that
bending moments of free and fixed head pile
converge at a common depth and become
practically identical beyond a certain distance
from the ground surface. This depth should be
correlated to the "active pile length" beyond
which a head-loaded pile behaves like an
infinitely long beam. Indeed, utilizing the
analytical relationship (Poulos and Davies,
1980, Randolph, 1981):
Ep

Es

La 1.5

statements (Dobry & O'Rourke, 1983,


Kavvadas and Gazetas, 1993) that higher
modes can produce larger kinematic bending.
However, the above results simply indicate a
possible cause when large kinematic induced
bending moments are observed and cannot
therefore be easily generalized. Certainly a
thorough investigation is required, including
various soil-pile configurations, in order to
establish further the effect of the soil
fundamental frequencies on the pile response.
EVALUATION OF PILE HEAD IMPEDANCE
FUNCTIONS

(3)

The evaluation of the pile head impedance


functions constitutes one of the essential
analysis steps for the inertial part of the
decomposition approach that was described
above. The soil-pile system was therefore
reanalyzed under pilehead harmonic loading, in
order to compare the numerically obtained
dynamic stiffness of the pile with analytical
expressions that have been proposed for the
case of a free head pile embedded in a
homogeneous soil. A convenient expression of
the dynamic impedance functions associated
with each vibration motion of the foundation is
given by the equation (Gazetas, 1987):

results in an active pile length equal to 8.6D,


which is very close to the depth below which
the distribution of the pile bending moments is
not affected by the different pilehead conditions.
Moreover, the bending moment developed on
the free head pile at the first natural frequency
of the soil deposit increases with depth,
reaching its maximum approximately at the midlength of the pile, while the bending moment of
a fixed head pile is a monotonically decreasing
function of depth.
Similar findings are reported in the study of
Nikolaou et al., 2001 where a Beam on
Dynamic
Winkler
Foundation
modeling
approach was implemented. It is interesting to
note though that based on the 3D FE analyses
obtained herein, higher deformation modes
generated larger pile bending moments for both
free and fixed head pile. This is in contrast to
the conclusions drawn by Nikolaou et al., 2001
and Gazetas et al. 1992, where the significant
overall drift of the first mode was associated to
larger bending moments. Although the pile
curvature is affected by the overall drift between
the top and the bottom of the pile, which is
indeed larger in the first natural frequency of the
soil, the curvature imposed on the pile due to
the "wavy" shape of the higher modes seems to
be the dominant response mechanism in this
case. Consequently, the variation of the
kinematic bending moment with frequency
doesnt follow the pattern of the free field soil
response shown in Fig 1 where greater
amplification is observed at the first
fundamental frequency of the soil deposit. This
observation is in agreement with some earlier

K() = K() + i C()

(4)

where the real part of the complex valued


impedance, K() , represents the stiffness and
the inertia of the supporting soil while the
component C() (imaginary part) reflects
the radiation and the material damping of the
system. Neglecting the vertical component of
motion, the dynamic stiffness of a free head pile
can be expressed in a matrix form as:
XX ()

XR ()

K RX ()

K RR ()

K() =

() + i C XX

K() =

RX () + i C RX

(5)

XR () + i C XR

RR () + i C RR

where the indices XX, XR and RR correspond


to the swaying, the cross swaying rocking and
the rocking component of the foundation motion
respectively. The pilehead dynamic stiffness

348

matrix given in Equation 5, was obtain herein by


the inversion of the corresponding flexibility
matrix that was in turn constructed from the
numerically derived pilehead displacements
and rotations due to a steady state horizontal
force or moment imposed on the pilehead. Fig 4
shows the pilehead impedance functions for the
free head pile, which are in a satisfactory
agreement with the analytical expressions
proposed by Gazetas et al., 1992. Indeed the
damping coefficient is reduced with increasing
frequency in each vibration mode of the
foundation while the real part of the dynamic
soil stiffness, K() , is practically insensitive to
the frequency content of the excitation force.
However, a slight difference is observed
regarding the amplitude of the real part of the
impedance functions, where the FE analyses
result in lower stiffness values with respect to
the analytical expressions, especially for the
swaying and the coupled swaying-rocking
motion. This difference that occurs also for the
static case (f=0) and is subsequently
reproduced with increasing frequency, should
be attributed to the different modeling
approximations introduced in each approach.
The latter is further established in Table 1,
given in the following page, where the pile
horizontal static stiffness obtained from the FE
analysis is compared to the well-known
relationship (Poulos and Davis, 1980):

14

34

(6)
K s = 4 E p Ip
k
implementing different expressions that have
been proposed for the determination of the
modulus of the soil subgrade reaction:

k Es

(7)

where the coefficient depends mainly on the


soil profile, the type of the head loading and the
relative soil-pile stiffness. It is interesting to note
the computed range for the pile static stiffness
value, Ks, as a result of the different approach
utilized by the various researchers to obtain .
Given the aforementioned variation in predicting
the pile stiffness based on the analytical
expressions utilized, the FE analyses results
are considered satisfactory, thus verifying the
ability of the FE model to adequately reproduce
soil-pile interaction under pilehead loading.
In a following stage, the calculated
frequency dependent springs and dashpots are
introduced at the base of the structure, in order
to
incorporate
soil
compliance
and
consequently evaluate the seismic response of
the superstructure due to inertial interaction
effects. This final step of the decomposition
approach is not further examined herein.
ANALYSIS OF THE COUPLED SOIL-PILESTRUCTURE SYSTEM

Although the decoupling of the total soil-pile-

600
400
200

f (Hz)

0
0

800

1500

600

1000

400
200

f (Hz)

0
10

10
5
f (Hz)

0
0

15
10
5
f (Hz)

0
10

10

f (Hz)
0

CRR (MNsecm/rad)

15

500
0

10

20

CXR (MNsecm/m)

20

CXX (MNsec/m)

KRR (MNm/rad)

800

2000

1000

Gazetas et al.
FE analysis

KXR (MNm/m)

KXX (MN/m)

1000

20
15
10
5
f (Hz)

0
0

Fig 4: Numerical results compared to analytical expressions regarding the pilehead impedance
functions of a free head pile

349

10

10

Table 1. Pile head static stiffness


Reference
Dobry et al, 1982

Proposed expression for


Ep
= 1.67
Es

Makris and
Gazetas, 1992
Kavvadas and
Gazetas (1993)

k (N/m2)

Ks (KN/m)

247000

521400

229830

494000

350000

670000

318400

630000

-0.053

= 1.29

1.2

3
1 - s2

18
E
s
Ep

18
L
= 2
D
1/12

Vesic, 1961

0, 65 Es D 4

1 - vs E p Ip

Davies and
Budhu, 1986

= 0.83

f uH =

*Gazetas et al.,
1992

K HH

1.3 K 0.18

358000

Es Deq

Ep
= D ES
E
s

0.21

532740

FE results (present study)

471000

*Compared to the FE analysis results shown in Fig.4

superstructure as a single degree of freedom


oscillator. The coupled system was also
analyzed in the frequency domain under
harmonic displacement applied at the bedrock
level. Emphasis was given on the effect of the
dynamic characteristics of the superstructure.
Three SDOF structures having different
fundamental frequencies were therefore
analyzed. The fixed base frequency, fstr, of each
structure was selected equal to the first, second
and third natural frequency of the foundation
soil respectively.
Table 2 summarizes the specific parameters
utilized in the analysis of the coupled system. It
is mentioned that the fundamental frequency of

structure system provides a clear identification


of the separate role of inertial and kinematic
interaction on the final seismic response of the
superstructure, the implementation of the
coupled system provides on the other hand a
direct and probably more realistic estimation of
the pile and structural response, since inertial
and kinematic effects are simultaneously
accounted for. The dynamic response of the
coupled soil-pile-structure system was therefore
investigated in a second stage of analysis,
implementing a single pile supported structure
founded on a homogeneous soil layer (Fig 5a).
For this reason a mass was attached to an
extension of the pile shaft, thus modeling the
20

10
Str.1
Str.2

12

Str.3

8
4

SSI

6
4
2

0
0 1 2 3 4 5 6 7 8 9 10
f(Hz)

(a)

Fixed Base
8
f (Hz)

Ampitude

16

(b)

0
Str.1

Str.2

Str.3

(c)

Fig 5: a) The coupled soil-pile-structure system studied b) Structural top to base displacement ratio c)
Effect of SSI on the fundamental period of the structure

350

Table 2. Investigated cases of soil-pile-structure interaction


Soil pile properties

Structural properties

H(m)

Vs

Dp

L/D

Ep/Es

Structure

m(t)

h(m)

Dstr.

EIstr./EIp

fstr
(Hz)

30

200

1.5

20

1000

Str.1

100

10

0.75

0.065

1.6

30

200

1.5

20

1000

Str.2

100

10

1.35

0.66

30

200

1.5

20

1000

Str.3

100

10

1.75

1.85

8.4

the superstructure was obtained by modifying


the stiffness of the superstructure. Since the
mass as well as the height of the superstructure
have been verified to play an important role in
the interaction mechanism (Kirtas et al. 2007),
they were kept constant throughout the
parametric investigation in order to reduce the
factors that contribute to the modification of the
fundamental frequency of the structure due to
soil structure interaction. The following
paragraphs present typical results obtained
from the numerical analyses of the coupled
system, emphasizing on the role of soil-pilestructure interaction on the pile and structural
dynamic response.

foundation. The comparative distributions of the


steady state bending moment amplitudes along
the pile length presented in Fig 6a, 6b and 6c
correspond to the first, second and third
fundamental frequency of the soil deposit
respectively after being normalized to the
maximum bending moment that occurred at
each frequency. More specifically, the bending
moment distributions depicted in Fig 6a are
normalized to the maximum bending moment
observed at the first natural frequency, while
the bending moment distributions depicted in
Fig 6b and 6c are normalized to the maximum
bending moment developed at the second and
third natural frequency of the soil deposit
respectively. The pile bending moments
obtained at the same frequencies considering
only kinematic interaction are also depicted in
Fig 6. It is observed that the bending moments
generated due to the combined effect of inertial
and kinematic interaction converge at a certain
depth beyond which they become practically
identical to the bending moments developed
due to solely kinematic interaction. Therefore,
structural oscillations affect pile response near
the pile head where the inertial interaction is
expected to become significant especially when
the structure is in or near resonance. The latter
is more pronounced in the case of the flexible
superstructure Str.1. Soil structure interaction
had a minor effect on the fundamental
frequency of the particular structure and
consequently the SSI frequency remained close
to the fixed base case, which in turn was
considered equal to the first natural frequency
of the soil deposit. Thus, the large bending
moment observed at the pilehead of the Str.1
pile foundation (Fig.6a) is actually the result of a
double resonance phenomenon where both the
soil and the structure are in resonance. A
similar pile response is observed in the case of

Effect of soil-pile-structure interaction on the


structural fundamental frequency

The increase of the fundamental period of


structures constitutes one of the main effects of
soil-structure interaction that has been widely
verified (Wolf, 1985, Gazetas, 1987). In order to
quantify this effect for the particular soil-pilestructure systems examined herein, the
structural top to base displacement ratio was
computed in each case (Fig 5b). The effect of
soil-structure interaction on the fundamental
frequency of the structure is depicted in Fig.5c,
where the effective (SSI) fundamental
frequency of the superstructure is compared to
the fixed base frequency for each examined
case. It is observed that stiffer structures (e.g
Str.3) are subjected to stronger SSI effects,
resulting therefore in a significant decrease in
the fundamental frequency of the structure.
Effect of soil-pile-structure interaction on the pile
bending moments

Soil-pile-structure interaction effects were


subsequently investigated in terms of the pile
bending moments developed on the pile

351

0.5

M / Mm ax
1

1.5

M / Mm ax
0.5
1

1.5

10

10

10

z(m)

15
20

Kinematic-Free Head
Kinem.+Inert. (Str.1)

25

Kinem.+Inert. (Str.2)
Kinem.+Inert. (Str.3)

30

Kinematic-Fixed Head

z ( m)

z(m)

15

20

25

25

30

30

(b)

1.5

15

20

(a)

M / Mm ax
0.5
1

(c)

Fig 6: Distribution of bending moment amplitude at the (a) first (b) second and (c) third natural
frequency of the soil deposit.

distress at greater depths. The latter is further


illustrated in Fig 7a where the frequency
variation of the depth that the maximum steady
state pile bending moment occurs is depicted. It
is interesting to note that the frequency range
where the maximum bending moment occurs at
the pilehead is increasing with the fundamental
frequency of the superstructure. However,
beyond a certain frequency the location of the
maximum pile bending moment is shifted away
from the pilehead, indicating additional cross
sections along the pile length that may be
subjected to large bending moments especially
under high frequency excitations. Similar
observations regarding the interplay between
the kinematic and inertial interaction in the
development of pile forces under seismic
loading have also been reported by Kaynia and
Mahzooni, 1996, where the dynamic response
of a SDOF oscillator founded on a typical 5x5
pile group was examined implementing a threedimensional
Green's
function-based
formulation.

the stiffer structure Str.3, where in this case the


effective (SSI) frequency of the structure is
close to the second natural frequency of the soil
deposit. Therefore, the distribution of the
bending moments that corresponds to the pile
foundation of Str.3 (Fig 6b), where the largest
bending moment is also observed at the
pilehead (a clearly inertial effect), results from
the interplay between the soil in resonance and
the structure near resonance. On the other
hand, the pile bending moment profiles of Fig
6c obtained at the third natural frequency of the
soil depict the predominant role of kinematic
effects on the pile bending moments when the
structure responds at a frequency different from
that of the imposed motion while the soil is in
resonance. Conclusively, it can be stated that
inertial interaction prevails when the excitation
frequency is close to the effective (SSI)
frequency of the superstructure imposing large
bending moments at or near the pilehead. This
should be correlated with the single pile
foundation employed herein, where the large
inertial force that is developed on the
superstructure due to resonance effects is
directly transmitted as a shear force and
bending moment onto the pilehead (Mylonakis
et al., 1997). On the other hand, when the
fundamental frequency of the input motion
excites higher soil modes and the structure is
not in resonance, strong kinematic effects are
activated that may also generate significant pile

Effect of soil-pile-structure interaction on the pile


bending moments

The response of the coupled soil-pilestructure system was further investigated in


terms of the pilehead motion U'p (Fig 8a). For
this reason the pile head to free field
displacement ratio (I'u=U'p/Uff) was computed for
each one of the examined cases. The obtained
results are shown in Fig 7b where the kinematic

352

3.5

3
2.5

15
Iu , I'u

z(m)

10

20
25

2
1.5
4

30
0

9 10

0.5

Kinematic_Free Head

Kinem.+Inert. (Str.1)

Kinem.+Inert. (Str.2)

Kinem.+Inert. (Str.3)

f(Hz)

Kinematic_Fixed Head

4 5 6
f(Hz)

9 10

(b)

(a)

Fig 7: a) Depth at which maximum bending moment occurs along the pile as a function of the
excitation frequency b) Pile head to free field displacement ratio

field response. The particular variation pattern


is in agreement with the analyses of real
earthquake events that have been recorded on
a pile-supported structure (Ohta et al., 1980).
Trying to identify the specific frequencies it was
observed that the frequency, where the
deamplification of the pilehead motion occurs,
coincides with the modified fundamental
frequency of the superstructure due to soilstructure interaction effects (Fig 5b). On the
other hand, the particular frequencies where the
pilehead motion is amplified with respect to the
free field motion were found to correspond to
the fundamental frequency of the pile-structure
system, that was in turn obtained from the
structural top to free field displacement ratio
(Uss/Uff) (Fig.8b). This ratio results actually
From the structural top to soil base motion ratio
(Uss/Ug) after being divided by the soil

interaction factors for free and fixed pile head


conditions are also plotted. It should be
mentioned though that the pile head to free field
motion ratio that is derived from the analysis of
the coupled system is not directly correlated
with the kinematic interaction factors, which are
conventionally obtained considering a massless
superstructure. However, the investigation of
the pilehead motion under the combined effect
of inertial and kinematic interaction indicated
the natural frequencies of the system that
actually affected the superstructure's base
motion. Indeed the frequency variation of I'u
seems to be dominated by two discrete
frequencies; a lower frequency (marked in Fig
7b with circles) where the pilehead motion is
amplified and a higher one (marked in Fig 7b
with rectangles) where the pilehead motion is
suddenly deamplified with respect to the free

20
Str.1

Ampitude

16

Str.2

12

Str.3

8
4
0

0 1 2 3 4 5 6 7 8 9 10
f(Hz)

(b)

(a)

Fig 8: a) Illustration of the response parameters utilised in the investigation of the pilehead motion b)
Structural top to free field displacement ratio

353

amplification ratio (Uff/Ug), thus removing the


effect of the soil response. Furthermore, it is
observed that the effect of the coupled
interaction mechanism on the pilehead
horizontal motion becomes more pronounced
as the frequency of the superstructure
increases, resulting in a stronger amplification
at the fundamental frequency of the pilestructure system.

ANSYS Inc (2000) ANSYS Users manual. Version


8.1, SAS IP, Houston, USA
Berill J., Yasuda S., "Liquefaction and piled
foundations:
Some
issues,"
Journal
of
earthquake engineering, Vol. 6,Special Issue 1,
pp 1-41, 2002
Davies T.G., and Budhu M., "Non-linear analysis of
laterally loaded piles in heavily overconsolidated
clays," Geotechnique, 36(4), 527-538,1986

CONCLUSIONS

Dobry R., Vicente E., O'Rourke T. and Roesset J.,


"Horizontal stiffness and damping of single piles",
Journal of the Geotechnical Engineering Division,
ASCE, Vol.108, No.GT3, 1982

Certain aspects of the soil-pile-structure


interaction mechanism were discussed in this
paper in terms of the kinematic soil-pile
interaction effects, the pile head impedance
functions and the dynamic response of the
coupled system. The harmonic response
analyses that were performed on a validated
finite element model aimed at obtaining an
insight on the effect of the natural frequencies
of the system on pile and structural dynamic
response. Based on the results of the numerical
analyses obtained herein it was observed that
significant kinematic induced bending moments
are developed on the pile at greater depths,
when higher modes of the surrounding soil are
excited. Thus, the maximum pile bending
moment due to kinematic interaction may for
certain soil-pile cases occur at frequencies
higher than the fundamental frequency of the
soil deposit. The examination of the entire soilpile-superstructure system showed on the other
hand that structural vibrations have a
paramount effect on the pile bending moments
when the frequency of the seismic motion is
close to the effective (SSI) fundamental
frequency of the structure, resulting in large
bending distress at the pilehead. Finally, the
combined effect of inertial and kinematic
interaction resulted in a significant amplification
of the pilehead motion with respect to the free
field response, which was observed at the
fundamental frequency of the pile-structure
system.

Dobry R. & O'Rourke M.J., Discussion on "Seismic


response of end-bearing piles" by FloresBerrones R & Whitman RV, Journal of
Geotecnhical Engineering Division, ASCE, p.109,
1983
CEN, European Committee for Standardisation.
prEN 19985: Eurocode 8: Design of Structures
for Earthquake Resistance, Part 1: General rules,
seismic actions and rules for buildings, Brussels,
2002
Fan K., Gazetas G., Kaynia A., Kausel E. and
Ahmad S., "Kinematic seismic response of single
piles and pile groups" Journal of Geotechnical
Engineering, ASCE, 117(12), 557-574, 1991
Fan K. and Gazetas G., "Seismic response of single
piles and pile groups" National Center for
Earthquake Engineering Research, Technical
Report NCEER-91-003, 1991
Gazetas G., "Seismic response of end bearing
single piles" Soil Dynamics and Earthquake
Engineering, 3(2), 82-93, 1984
Gazetas G., "Simple physical methods for
foundation impedances", in Dynamic Behaviour
of Foundations and Buried Structures, Banerjee
P. and Butterfield R. (eds), Elsevier Applied
Science, 45-93, 1987
Gazetas G., Fan K., Tazoh T., Shimizu K., Kavvadas
M. and Makris N., "Seismic pile group structure
interaction" Piles Under Dynamic Loads,
Geotechnical Special Publication No.34, ASCE,
S.Prakash (ed.), 56-93, 1992
Gazetas G. & Mylonakis G., "Seismic soil-structure
interaction: New evidence and emerging issues"
Geotechnical Earthquake Engineering and Soil
Dynamics, Vol.III, Dakoulas P. Yegian M and
Holtz R (eds), Geotechnical Special Publication,
No.75, ASCE, 1119-1174, 1998

ACKNOWLEDGEMENTS

The first author would also like to acknowledge


the contribution of IKY foundation.
REFERENCES

354

Engineering and structural Dynamics, 26, 337359, 1997

Guin J. and Banerjee P.K., "Coupled soil-pilestructure interaction analysis under seismic
excitation" Journal of Structural Engineering,
124(4), 1998

Mylonakis G., "Simplified model for seismic pile


bending at soil layer interfaces", Soils and
Foundations, 41(4), 47-58, 2001

Kagawa T. and Kraft L.M., "Lateral pile response


during earthquakes," Journal of the Geotechnical
Engineering Division, ASCE, 108(4), 554-569,
1982

Nikolaou S., Mylonakis G., Gazetas G. & Tazoh T.,


"Kinematic pile bending during earthquakes:
analysis and field measurements", Geotechnique,
51(5), 425-440, 2001

Kavvadas M. and Gazetas G., "Kinematic seismic


response and bending of free-head piles in
layered soil," Geotechnique, 43(2), 207-222,
1993

Nogami T., "Dynamic group effect in axial responses


of grouped piles" Journal of the Geotechnical
Engineering Division, ASCE, 109(7), 225-243,
1983

Kaynia A.M. and Novak M., "Response of pile


foundations to Rayleigh waves and to obliquely
incident body waves, " Earthquake Engineering
and Structural Dynamics, 21(4), 303-318, 1992

Ohta T., Uchiyama S., Niwa M. and Ueno K.,


"Earthquake response characteristics of structure
with pile foundation on soft subsoil layer and its
simulation analysis," Proceedings 7th World
Conference on Earthquake Engineering, Istanbul,
Turkey, Vol.3, pp 403-410, 1980

Kaynia A.M. and Mahzooni S., "Forces in pile


foundations under seismic loading," Journal of
Engineering mechanics, 122(1), 46-53, 1996
Kirtas E., Trevlopoulos K., Rovithis E., Pitilakis K.,
"Discussion of the fundamental period of fixed
base systems including soil-structure interaction",
4th International Conference on Earthquake
Geotechnical Engineering, Thessaloniki, Greece,
2007 (submitted)
Lok T.M., Pestana J. and Seed R.B., "Numerical
modeling and simulation of coupled seismic soilpile-structure
interaction"
Geotechnical
Earthquake Engineering and Soil Dynamics,
Vol.III, Dakoulas P. Yegian M and Holtz R (eds),
Geotechnical Special Publication, No.75, ASCE,
1211-1222, 1998

Poulos H.G. and Davis E.H., Pile foundation analysis


and design, John Wiley and Sons, New York,
N.Y, 1980.
Randolph M.F., "The response of flexible piles to
lateral loading," Geotechnique, 31(2), 247-259,
1981
Roesset J.M., "Soil amplification of earthquakes" in
Numerical Methods in Geotechnical Engineering,
Desai C.S., and Christian J.T. (eds), McGraw-Hill,
1977
Tazoh T., Shimizu K. & Wakahara T., "Seismic
observations and analysis of grouped piles". In
Dynamic response of pile foundations:
experiment,
analysis
and
observation,
Geotechnical Special Publication No.11, ASCE,
1987

Makris N. and Gazetas G. "Dynamic pile-soil-pile


interaction, part II: Lateral and Seismic response"
Earthquake
Engineering
and
Structural
Dynamics, 20(2), 145-162, 1992

Tokimatsu K.H., Oh-Oka K., Satake Y., Shamoto Y.,


Asaka, "Effects of Lateral Ground Movements
on Failure Patterns of Piles in the 1995
Hyogoken-Nambu Earthquake" Geotechnical
Earthquake Engineering and Soil Dynamics III,
Geotechnical Special Publication 75, ASCE, Vol.
2, pp. 1175-1185, 1998

Makris N., Badoni D., Delis E., Gazetas G.,


"Prediction of observed bridge response with
soil-pile-structure
interaction"
Journal
of
Structural Engineering, 120(10), 2992-3011,
1994
Meymand P.J., "Shaking table scale model tests of
nonlinear soil-pile-superstructure interaction in
soft clay" PhD Thesis, University of California,
Berkeley, 1998

Wolf J.P., Dynamic Soil-Structure interaction,


Prentice Hall, Englewood Cliffs, NJ, 1985
Wilson D.W., "Soil-pile-superstructure interaction in
liquefying sand and soft clay, PhD Thesis,
Department of Civil and Environmental
Engineering, University of California, Davis, 1998

Mizuno H., " Pile Damage during Earthquake in


Japan (1923-1983)" Dynamic Response of Pile
Foundations, Geotechnical Special Publication,
No. 11, ASCE, pp. 53-78, 1987

Vesic A.S., "Beam on elastic subgrade and the


Winkler hypothesis," Proc., 5th International
Conference on Soil Mechanics and Foundation.
Engineering, Paris, Vol. 1, 845-850, 1961eering
Division, ASCE, 108(4), 554-569, 1982

Mylonakis G., Nikolaou S. and Gazetas G., "Soilpile-Bridge seismic interaction: Kinematic and
inertial effects. Part I: Soft soil" Earthquake

355

Kavvadas M. and Gazetas G., "Kinematic seismic


response and bending of free-head piles in
layered soil," Geotechnique, 43(2), 207-222,
1993

Nogami T., "Dynamic group effect in axial responses


of grouped piles" Journal of the Geotechnical
Engineering Division, ASCE, 109(7), 225-243,
1983

Kaynia A.M. and Novak M., "Response of pile


foundations to Rayleigh waves and to obliquely
incident body waves, " Earthquake Engineering
and Structural Dynamics, 21(4), 303-318, 1992

Ohta T., Uchiyama S., Niwa M. and Ueno K.,


"Earthquake response characteristics of structure
with pile foundation on soft subsoil layer and its
simulation analysis," Proceedings 7th World
Conference on Earthquake Engineering, Istanbul,
Turkey, Vol.3, pp 403-410, 1980

Kaynia A.M. and Mahzooni S., "Forces in pile


foundations under seismic loading," Journal of
Engineering mechanics, 122(1), 46-53, 1996
Kirtas E., Trevlopoulos K., Rovithis E., Pitilakis K.,
"Discussion of the fundamental period of fixed
base systems including soil-structure interaction",
4th International Conference on Earthquake
Geotechnical Engineering, Thessaloniki, Greece,
2007 (submitted)
Lok T.M., Pestana J. and Seed R.B., "Numerical
modeling and simulation of coupled seismic soilpile-structure
interaction"
Geotechnical
Earthquake Engineering and Soil Dynamics,
Vol.III, Dakoulas P. Yegian M and Holtz R (eds),
Geotechnical Special Publication, No.75, ASCE,
1211-1222, 1998

Poulos H.G. and Davis E.H., Pile foundation analysis


and design, John Wiley and Sons, New York,
N.Y, 1980.
Randolph M.F., "The response of flexible piles to
lateral loading," Geotechnique, 31(2), 247-259,
1981
Roesset J.M., "Soil amplification of earthquakes" in
Numerical Methods in Geotechnical Engineering,
Desai C.S., and Christian J.T. (eds), McGraw-Hill,
1977
Tazoh T., Shimizu K. & Wakahara T., "Seismic
observations and analysis of grouped piles". In
Dynamic response of pile foundations:
experiment,
analysis
and
observation,
Geotechnical Special Publication No.11, ASCE,
1987

Makris N. and Gazetas G. "Dynamic pile-soil-pile


interaction, part II: Lateral and Seismic response"
Earthquake
Engineering
and
Structural
Dynamics, 20(2), 145-162, 1992

Tokimatsu K.H., Oh-Oka K., Satake Y., Shamoto Y.,


Asaka, "Effects of Lateral Ground Movements
on Failure Patterns of Piles in the 1995
Hyogoken-Nambu Earthquake" Geotechnical
Earthquake Engineering and Soil Dynamics III,
Geotechnical Special Publication 75, ASCE, Vol.
2, pp. 1175-1185, 1998

Makris N., Badoni D., Delis E., Gazetas G.,


"Prediction of observed bridge response with
soil-pile-structure
interaction"
Journal
of
Structural Engineering, 120(10), 2992-3011,
1994

Wolf J.P., Dynamic Soil-Structure interaction,


Prentice Hall, Englewood Cliffs, NJ, 1985

Meymand P.J., "Shaking table scale model tests of


nonlinear soil-pile-superstructure interaction in
soft clay" PhD Thesis, University of California,
Berkeley, 1998

Wilson D.W., "Soil-pile-superstructure interaction in


liquefying sand and soft clay, PhD Thesis,
Department of Civil and Environmental
Engineering, University of California, Davis, 1998

Mizuno H., " Pile Damage during Earthquake in


Japan (1923-1983)" Dynamic Response of Pile
Foundations, Geotechnical Special Publication,
No. 11, ASCE, pp. 53-78, 1987

Vesic A.S., "Beam on elastic subgrade and the


Winkler hypothesis," Proc., 5th International
Conference on Soil Mechanics and Foundation.
Engineering, Paris, Vol. 1, 845-850, 1961

Mylonakis G., Nikolaou S. and Gazetas G., "Soilpile-Bridge seismic interaction: Kinematic and
inertial effects. Part I: Soft soil" Earthquake
Engineering and structural Dynamics, 26, 337359, 1997
Mylonakis G., "Simplified model for seismic pile
bending at soil layer interfaces", Soils and
Foundations, 41(4), 47-58, 2001
Nikolaou S., Mylonakis G., Gazetas G. & Tazoh T.,
"Kinematic pile bending during earthquakes:
analysis and field measurements", Geotechnique,
51(5), 425-440, 2001

356

Utilization of P-Y Curves for Estimating Soil Pile Interaction


under Seismic Loading

E. Rovithis1, E. Kirtas1, K. Pitilakis1


1

Department of Civil Engineering, Aristotle University of Thessaloniki, Greece

Abstract
To predict the seismic response of pile-supported structures, the p-y method has been
extensively used, where the soil is represented as a series of independent springs
distributed along the pile shaft. Along these lines several p-y curves for different soil-pile
systems have been proposed which are mainly based on in-situ pile tests under static or low
frequency cyclic loading conditions. However, soil-pile interaction under seismic excitation
becomes more complex due to the incident seismic waves scattered by the pile, thus
modifying the p-y relationship and introducing an additional damping mechanism at the soilpile interface. In this paper, dynamic soil-pile interaction is estimated based on backcalculated p-y curves, disregarding in a first stage any potential soil-pile gapping
mechanism. Pile displacements and soil reactions are derived through double integrating
and differentiating respectively the bending moments obtained along the pile shaft. The p-y
curves generated at each depth are utilized to derive frequency dependent springs and
dashpots, which may then be implemented within the framework of a Beam on Dynamic
Winkler foundation modeling of the soil-pile system. The proposed procedure is validated
through centrifuge tests results of a coupled soil-pile-structure system under real
earthquake excitation and is also compared to existing analytical formulas of frequency
dependent springs and dashpots under steady state harmonic excitation applied on the pile
head. Analysis results reveal that for each one of the loading scenarios considered, soil-pile
interaction mechanism is adequately captured while utilizing existing analytical expressions
for the computation of dynamic soil spring supports may under certain conditions lead to an
overestimation of pile and structural response when the coupled soil-pile-structure system is
analyzed.

INTRODUCTION

reactions may be accounted for, the p-y curves


obtained under monotonic or cyclic loading
conditions do not directly describe neither the
reduction of the soil stiffness with the increasing
amplitude of the seismic motion (more
pronounced for soft soil conditions) nor the soil
inertia and radiation damping effects which play
a fundamental role in the soil-pile interaction
mechanism (Angelides and Roesset, 1981).
Consequently, the implementation of a Winkler
model for analyzing soil-pile interaction under
strong ground motion needs to take into
consideration the modification of the p-y
relationship due to the incident seismic waves
that are scattered by the pile. Furthermore the
presence of hysteretic and radiation damping
affects soil-pile interaction, introducing the

Several procedures have been developed


during the last two decades in order to
determine the dynamic response of piles
subjected to horizontal or vertical loads. Most of
these procedures adopt the Winkler approach
where the pile is modeled as a beam on elastic
foundation and the soil is represented as a
series of independent springs attached to the
pile shaft. The determination of the spring
stiffness is usually based on certain p-y curves
that have been established for particular soil
types (Matlock, 1970, Reese et al. 1974, Reese
and Welch, 1975, Janoyan et al. 2001).
Although the non linear relationship between
the pile displacements and the lateral soil

357

complex valued soil stiffness as a function of


the frequency content of the seismic motion.
In order to incorporate these dynamic
effects in the framework of a Winkler
approximation, simplified expressions of
frequency dependent springs and dashpots
(Dobry et al., 1982, Gazetas and Dobry, 1984,
Kavvadas and Gazetas, 1993) as well as more
rigorous analytical p-y models (Boulanger et al.,
1999, Lok et al. 1998) have been proposed for
seismic soil-pile interaction analysis. However,
little research has so far taken place towards
the correlation of the dynamic stiffness at the
soil-pile interface with the p-y loops that are
generated under seismic loading. Nevertheless,
the fundamental properties of the dynamic p-y
loops such as the slope and the area of the
loop provide a straightforward estimation of the
springs and dashpots that could be utilized to
model dynamic soil-pile interaction.
In this paper a generalized procedure is
presented where dynamic soil-pile interaction is
evaluated based on back-calculated p-y loops,
disregarding in a first stage any potential soilpile gapping mechanism. Pile displacements
and lateral soil reactions are initially computed
separately, based on the bending moments
developed along the pile shaft. Frequency
dependent springs and dashpots are then
estimated utilizing the basic characteristics of
the respective p-y loops obtained at each
depth.
The adopted procedure is validated through
centrifuge tests results of a coupled soil-pilesuperstructure system under real earthquake
excitation introduced at the base of the soil
profile. Due to the consideration of the coupled
system, the effect of both inertial and kinematic
interaction on the developed pile bending
moments is therefore accounted for. The
procedure was also applied to a single pile
embedded in a uniform soil profile and the
obtained results were compared to existing
analytical formulas of frequency dependent
springs and dashpots that have been proposed
for pilehead harmonic loading. Finally, a
comparative analysis is performed where the
effect of the springs and dashpots coefficients
on the seismic response of the coupled system
is investigated.

two discrete parts of analysis. The first


concerns the derivation of the dynamic p-y
loops from the bending moments generated
along the pile shaft, while the second one deals
with the estimation of frequency dependent
springs and dashpots based on the p-y loops
derived in the previous stage. Therefore, the
main scope of the procedure is to obtain
stiffness and damping components of the
dynamic soil-pile interaction mechanism that
could be used in a simplified beam on dynamic
Winkler foundation model of the soil-pile
system. Each one of the analysis steps is
explained further in the following paragraphs.
Derivation of p-y curves from recorded pile
bending moments

Under seismic excitation the magnitude of


the pile bending moment is generally a function
of time as well as of the vertical distance along
the pile shaft z. Therefore, for each time
instance, ti, a different bending moment
distribution M(z,ti) is generated along the pile.
Using simple beam theory, the seismic p-y
behaviour can be back calculated from the
bending moment distribution according to the
equations:
2 M(z, t i )
(1)
p(z, t i ) =
z 2

pile

(z, t i ) =

EI

M(z, ti )

(2)

where p is the lateral soil reaction on the pile,


ypile is the absolute lateral displacement of the
pile and EI is the flexural stiffness of the pile.
Having obtained the p and ypile distribution along
the pile at each time instance, the time histories
of the lateral soil reaction and the pile
displacement are then computed at each depth
z. The relative displacement between pile and
soil, y, is finally obtained by ypile after
subtracting the free field soil deflected shape
ysoil. Hence p-y loops are generated at a given
depth. It should be mentioned though that one
of the main tasks in estimating p(z,ti) and
ypile(z,ti) is the selection of a proper
interpolation function to describe pile bending
moment distribution.

OUTLINE OF THE ADOPTED PROCEDURE

Interpolation functions

The procedure adopted herein comprises of

Two of the most common interpolation


techniques usually employed to describe

358

bending moments distribution along the pile


involve fitting either cubic splines between
successive data points (Finn et al., 1983) or
higher order polynomial functions to all data
points (Ting, 1987). In the present study a 4th
order polynomial is adjusted on the bending
moments profile at each time instance ti. Thus
the bending moment distribution is given by the
equation:

pmax

ymax y
AD
Fig 1: Schematic illustration of the fundamental p-y
characteristics

M(z, ti) = A ti z 4 + Bti z3 + C ti z 2 + D ti z + E ti (3)

where Ati, Bti, Cti, Dti, Eti are the constant terms
of the polynomial. Inclusion of the quadratic
term in the polynomial function may result in
large lateral resistance values at the ground
surface (Wilson, 1998). Thus, the ability of the
adopted interpolation function to adequately
reproduce lateral soil reactions was further
checked with experimental results from a
statically loaded single pile obtained by
(Georgiadis et al. 1992). In the aforementioned
study, a fourth order spline function that
interpolates between successive experimental
points was employed. The satisfactory
agreement that was observed between the
different interpolation techniques (Rovithis et
al., 2007a) verified further the ability of the
polynomial approximation to describe the pile
response, justifying therefore its use during the
adopted procedure.

where AD is the area of the p-y loop which


corresponds to the sum of the hysteretic and
radiation damping, is the frequency of the
input motion and ymax corresponds to the
maximum displacement observed in the p-y
loop. The parameters AD and ymax are
schematically clarified in Fig.1.
The combination of equations 4 and 5 leads
to the computation of the real part kx() of the
complex valued impedance Sx:
2

p max A D

k x () =
-
y max y 2
max

The force (per unit length of the pile) p to


displacement y ratio of the Winkler medium at
every depth, defines the complex valued
frequency-dependent impedance (Kavvadas
and Gazetas, 1993):

AD
y 2max

(6)

VALIDATION OF THE SOIL-PILE


INTERACTION MODEL
Centrifuge testing of single pile supported
structures

(4)

Centrifuge testing has the advantage of


reproducing the initial stress field of soil models
of significant depth since it is conducted in high
gravitational environments. Moreover, pile
bending moments from centrifuge and/or
shaking table tests of soil-pile systems have
been widely utilized to derive experimental p-y
curves (Meymand, 1998, Wilson, 1998 among
others).
In this study, a well-documented series of
dynamic centrifuge tests of pile-supported
structures in soft ground was implemented for
the validation of the procedure described

where the real part kx() represents the


stiffness of the supporting soil and the
imaginary component Cx() reflects the
hysteretic and radiation damping generated in
the system. Having obtained the p-y loop at
each level along the pile length, the damping
coefficient is computed via the expression
(Badoni & Makris, 1995):

c x () =

where pmax is the maximum soil reaction


obtained by the double differentiating of the
bending moment. A similar approach was
adopted by Nogami et al. 1992, utilizing though
frequency independent parameters for the
description of the soil-pile interaction model that
was proposed.

Estimation of frequency dependent springs and


dashpots based on p-y curves

P
Sx = = k x () + i c x ()
y

1
2 2

(5)

359

process, the bending moments time histories


were initially band pass filtered and were then
utilized to obtain soil reactions and pile
displacements.

above. The tests were performed in a 9-mradius centrifuge at the University of California
at Davis (Wilson et al. 1997b, Boulanger et al,
1999). The models were tested in a flexible
shear beam (FSB) container at a centrifugal
acceleration of 30g. The double layer soil profile
comprises soft clay overlying dense sand. The
lower layer was a 11m thick, dense Nevada
sand layer with relative density Dr=75-80%. The
upper 6m thick layer was reconstituted Bay
Mud with significantly low undrained shear
strength, measured up to 8-10KPa at the level
of -6m.
The centrifuge test layout is schematically
shown in Fig 2a where the instrumentation of
the examined single pile supported structure
(SP1) is also depicted. The instrumentation of
the pile consists of six pairs of axial strain
gauges attached to the pile (five in the soft clay
layer and one in the dense sand layer).
Furthermore, accelerometers were placed at
the pilehead and the superstructure mass as
well as at different levels of the soil profile in
order to measure free field soil response.
The fixed base period of the single-pile
supported structure (SP1), which consisted of a
superstructure mass attached to an extension
of the pile, was measured equal to 0.29sec.
The input motion applied at the base of the
centrifuge apparatus was a strong motion
record from Port Island in the 1995 Kobe
earthquake scaled to PGA=0.05g and
PGA=0.20g respectively. Fig 2b shows the
bending moments recorded along the pile shaft
for the low amplitude motion. In order to avoid
high frequency noise during the differentiation

Derivation of p-y curves

Based on the recorded bending moments


for each intensity level of the base excitation, py curves were generated utilizing the procedure
described above. In order to obtain pile
displacements by double integrating bending
moments, two boundary conditions were
defined. The first one was determined at the
pile head, where the displacement time history
was calculated from the recorded acceleration,
while the second one was assigned to the pile
tip where the net displacement between soil
and pile was assumed to be zero. The latter
was based on the fact that due to the large
stiffness of the sand layer, pile deformations are
expected to follow the sand profile displaced
shape. Thus, the second boundary condition
was obtained by double integrating the soil
acceleration time history recorded at the
respective level. Typical results of the
computed p-y loops are shown in Fig 3 and Fig
4 corresponding to 0.05g and 0.2g amplitude of
the input motion respectively. The p-y loops
shown in Fig 3 and 4 were obtained at three
different levels along the pile shaft; at the depth
of 1.5m and 4.5m in the soft clay layer and at
the depth of 8.2m in the dense sand layer.
Reviewing the results it is observed that the
slope of the p-y loops is increasing with depth,
indicating higher soil stiffness levels, while on
the other hand the area of the p-y loop is

SP1

M (MNm)

Soft
clay

0.6
0.4
0.2
0
-0.2
-0.4
-0.6

6m

Dense
sand

sec
0

11m

MNm)

Bending/axial gauge

0.5
0

10

12

14

16

18

z=0.7m

z=1.5m

z=2.3m

z=3.8m

z=5.3m

z=8.4m

Accelerometer

(a)

(b)

Fig 2: a) Schematic of centrifuge layout and instrumentation b) Bending moment time histories
recorded along the pile shaft for the low amplitude motion (0.05g)

360

20

y50 = 2.5 D 50

decreasing corresponding to lower levels of


damping generated in the soil-pile system. The
above findings are related to the reduction of
the pile displacements with depth and the
associated linear response prevailing over the
nonlinear hysteretic action in the soil.
The computed p-y loops were also
compared to cyclic p-y curves (shown in Fig 3
and 4 with dashed gray lines) commonly used
in design practice. In the case of soft clay layer
the backbone p-y curve was based upon
Matlock's model (Matlock, 1970), which is
defined by the equation:
y
p
= 0.5
y
p ult
50

where D is the pile diameter, Np is the lateral


bearing capacity factor, is the average
buoyant unit weight, z is the depth, cu is the
undrained shear strength and 50 is the strain
corresponding to a stress of 50% of the ultimate
stress in a laboratory stress-strain curve. This
variable was taken equal to 0.005 based on
published laboratory test data for the particular
soil type (Boulanger et al., 1999) while the
value of parameter J was selected equal to 0.5
based on Matlock's recommendations for soft
clay.
For the underlying dense sand layer the p-y
curves proposed by Murchison and 'Neill
(1984), which are also adopted in the API
provisions (API, 1993) were implemented.
The comparison shows that commonly used
p-y curves can predict satisfactorily the secant
spring stiffness as long as the response
remains in the linear elastic range. For higher
levels (e.g 0.2g) of excitation though, where
significant hysteretic response is activated in
the soil, the cyclic p-y curves tend to
overestimate soil stiffness.

(7)

where pult is the ultimate lateral reaction and y50


is the lateral pile displacement at one-half of the
ultimate reaction, derived through the following
formulas:

pult = cu D N p

Np = 3 +

(9)
50

50

25

25

25

0
-25

P (KN/m)

50
P (KN/m)

P (KN/m)

(8)

' z J z
+
9
cu
D

0
-25

-50
-200 -100

y (mm)

100

200

0
-25

-50
-200 -100

y (mm)

(a)

(10)

100

200

-50
-200 -100

y (mm)

(b)

100

200

(c)

100

100

50

50

50

0
-50

P (KN/m)

100
P (KN/m)

P (KN/m)

Fig 3: p-y loops computed for 0.05g amplitude of the input motion at a) 1.5m b) 4.5m c) 8.2m

0
-50

-100
-200 -100

y (mm)

(a)

100

200

0
-50

-100
-200 -100

y (mm)

(b)

100

200

-100
-200 -100

y (mm)

100

(c)

Fig 4: p-y loops computed for 0.20g amplitude of the input motion at a) 1.5m b) 4.5m c) 8.2m

361

200

while each one of the pile nodes was connected


to a Kelvin element having spring and dashpot
properties as computed above.
Soil displacement time histories that were
calculated by double integrating the recorded
free field acceleration time histories, were the
imposed input motion at the ends of the Kelvin
elements. The numerical results obtained form
the analysis of the BDWF model were
compared to the recorded data in terms of pile
bending moments as well as pilehead and
superstructure acceleration time histories and
response spectra.
Typical results are shown in Fig.6 for the
two examined intensity levels of the input
motion. It is worth noting that, notwithstanding
the sharp increase in soil stiffness at the level of
the clay-sand interface, the peak bending
moment occurs close to the pilehead indicating
the effect of the inertial loading transmitted to
the pile due to the oscillation of the
superstructure (Mylonakis et al, 1997).
The very good agreement that is observed
between the numerical results and the
experimental recordings both in time and
frequency domain, proves the ability of the
BDWF model to adequately reproduce soil-pilestructure interaction effects. The procedure
adopted is therefore validated based on the
particular coupled system response under real
earthquake excitation. The application of the
procedure to different soil-pile systems under
different loading considerations as well is
examined in the following paragraph.

Analysis of the coupled system utilizing a BDWF


model

The experimental p-y loops obtained at


each location along the pile length are then
utilized to derive frequency dependent springs
and dashpots. The computation of the stiffness
and damping coefficients, kx() and cx(), was
performed using equations 5 and 6 where is
taken equal to the fundamental frequency of the
free field motion recorded at the respective
level. It should be mentioned that the Fourier
spectra peak amplitudes of the free filed motion
recorded in the soft clay layer appeared at the
frequency of 1Hz, which is also close to the
fundamental frequency of the superstructure
including the effect of soil structure interaction.
The utilization of the predominant frequency of
the input motion has also been adopted in
relevant soil-pile-structure interaction studies
(Makris et al. 1994) resulting in almost accurate
predictions of the system response. The
distribution of the real and imaginary part of the
impedance Sx (as defined in equation 4) is
shown in Fig 5a corresponding to the low
(0.05g) amplitude excitation.
Having obtained the distributed springs and
dashpots for each intensity level of the base
motion, the validation of the p-y procedure was
then performed utilizing a Beam on Dynamic
Winkler Foundation model of the coupled soilpile-structure system (Fig 5b). The seismic
response of the coupled system was analyzed
with the general purposed FE code ANSYS
(Ansys, 2000). The pile structure system was
modeled with linear elastic beam elements

0
0
2

z (m)

4000

KN/m
8000

12000

Superstructure

Soil response

Real part [K()]


Imag. part [C()]

kx()

Pile

cx()

10
12
14
16

Base motion

18

(a)

(b)

Fig 5: a) Distribution of the spring and dashpots along the pile for the low amplitude motion
b) Schematic illustration of the coupled system used for the validation of the p-y procedure

362

M(z, f) = Mo (z, f) ei[(2f)t+(f,z)]

COMPARISON WITH ANALYTICAL MODELS

(11)

Harmonic excitation at the pile head

0.2
0.1

Centrifuge
BDWF model

10

15

20

Sa(g)

(a)
0.6
0.3
0
-0.3
-0.6

tim e (sec)
0

10

(b)

15

0.8

0.6

0.4

10

0.2

12

14

2
T (sec)

(c)

BDWF model (Superstructure)

20

Mom ent (MNm )


0.5
1
1.5

Centrifuge (Superstructure)
0

tim e(sec)

BDWF model (Pilehead)

0.6
0.4
0

Acc. (g)

Centrifuge (Pilehead)

1
0.8

0.2

0
-0.1
-0.2

where the moment amplitude Mo(f,z) and the


phase angle (f,z), which are both functions of
the excitation frequency f and of the depth z,
are obtained from the numerical analyses of the
soil-pile system.
Fig 7b shows the distribution with depth of
the amplitudes of the bending moments
computed at the fundamental frequencies of the
soil deposit after being normalized to the
maximum bending moment that occurred at the
pile head. It is observed that bending moments
are practically insensitive to the excitation
frequency. Utilizing equation 11 the pile
bending moments can be computed in the time
domain for a given excitation frequency. Having
calculated the time histories of the pile bending
moments at each depth, the distribution of
moments along the pile's length can be
obtained at each time instance and may then
be utilized to derive the p-y loops and the
distributed springs and dashpots following the
procedure described above. The p-y loops
obtained at the pilehead that correspond to the
fundamental frequencies of the soil deposit are
shown in Fig 7c. It should be mentioned that the
results of Fig 7c are normalized to the
maximum p and y values observed in the first
natural frequency of the soil deposit. It is noted
that the shape of the loop resembles that of an

z(m)

Acc. (g)

In order to investigate further the


applicability of the p-y procedure, the case of a
single fixed-head, end bearing pile embedded
in a uniform soil profile was also examined
(Fig.7a). Therefore, a 30m thick homogeneous
soil layer over rigid bedrock was considered,
with shear wave velocity Vs=200m/sec, Poisson
ratio v=0.4 and hysteretic damping ratio
s=0.05. The first, second and third
fundamental frequency of the soil deposit
equals to 1.67Hz, 5Hz and 8.33Hz respectively.
The pile cross section was considered circular
while the relative soil-pile stiffness was taken
equal to Ep/Es=1000 and the pile slenderness,
i.e the ratio of the pile length L to the pile
diameter D, was selected equal to L/D=20. The
excitation motion introduced at the pile head
consists of a steady state horizontal force,
P=Poei2f*t, with amplitude Po=1KN and
frequency varying within the range of 1-10Hz.
The analysis of the soil-pile system was
performed in the frequency domain utilizing a
3D FE model. It should be noted that soil
response analyses were also performed in
order to verify that wave propagation effects are
adequately captured (Rovithis et al, 2007b).
Due to the particular type of the excitation
force, pile-bending moments are generally
described by the equation:

16

Clay - Sand
interface
Centrifuge (0.05g)
BDWF model (0.05g)
Centrifuge (0.20g)
BDWF model (0.20g)

(d)

Fig 6: Comparison of the BDWF model response to centrifuge recordings regarding a) Superstructure
acceleration (0.05g) b) Superstructure acceleration (0.20g) c) Response spectra (0.05g) d) Peak bending
moments

363

ellipse due to the harmonic type of loading that


was employed. Furthermore, when the pilehead
is subjected to higher excitation frequencies the
pile displacements are getting smaller and the
(secant) stiffness of the distributed springs
increases. This effect has also been observed
in Badoni & Makris, 1995 where the nonlinear
response of single piles is investigated. The
frequency dependent springs and dashpots
computed from the generated p-y loops,
utilizing equations 5 and 6, were then compared
to the following closed form expressions, which
are based on the 2D plane strain model of
Gazetas and Dobry, 1984:
k x 1.2 Es

in order to incorporate the generation of surface


waves due to the presence of the stress-free
ground surface.
Furthermore, the cx values obtained from
equations 13 and 15 apply for frequencies
higher than the first natural frequency, s, of the
soil deposit. For frequencies lower than s
radiation damping is significantly small
compared to the hysteretic type of soil damping.
On the other hand the expression for the spring
coefficient (Eq.12) has been determined by
matching the horizontal stiffness at the head of
the pile embedded in the Winkler soil model
and the continuum (finite element) soil model.
However, an approximation introduced in
obtaining equation 12 is to neglect the influence
of pile slenderness L/D and relative soil-pile
stiffness Ep/Es (Makris et al. 1996). This effect
is, in contrast, incorporated in the analytical
expression proposed by Kavvadas and
Gazetas, 1993, which is given by the following
equation corresponding to a pile embedded in a
two layered soil profile.

(12)

V
c x 2 a o-1 4 s Vs D 1 + La
V

5 4

+ 2 kx
s

(13)

The above formulas have also been used in


various soil-pile interaction studies (Makris and
Gazetas, 1992, Gazetas et al., 1992,
Mylonakis, 1995, Ahmad et al., 2006). The
parameter VLa represents the Lysmer's analog
velocity and is given by the equation:
3.4 Vs
V =
La (1 - v)

18

(14)

cx 2 a o-1 4

34

3
kx
1 - s2

k
+ 2 s x (15)

M/Mm ax

0.5

E
s
Ep

1 12

1 15

V
sa
V
sb

1.5

1.5

(16)

(17)

p/pmax

0.5

10

y/ymax

15

f=f1,soil

20

f=f2,soil

25

f=f3,soil

-1.5

-1

-0.5
0
-0.5

0.5

-1
-1.5

30

(a)

18
18

L
Es

z(m)

H
a
H
b

and which for the case of a homogeneous soil


and a pile of circular cross-section simplifies to:

At very shallow depths ( z 2.5D ), equation 13


is replaced by:

s Vs D
4

18
L

D

E D4
2

sa
kx
2 E I

1 - s p p

(b)

(c)

Fig 7: a) Examined soil-pile system b) Normalized distribution with depth of the amplitudes
of bending moments computed at the fundamental frequencies of the soil deposit
(c) Normalized pilehead p-y curves computed at the fundamental frequencies of the soil deposit

364

1.5

Since the examined coupled system consists of


a pile-supported structure founded on a twolayered soil deposit, the analytical expressions
given by equation 16 and equations 12 or 13
were employed as more appropriate to the
particular case. The soil shear wave velocity,
Vs, and the hysteretic damping ratio, s
introduced in the analytical equations, were
properly modified according to the developed
shear strains during the experiment. These
strain compatible values were determined from
an equivalent linear soil response analysis
which results in shear wave velocity profiles of
decreasing values with the intensity of the input
earthquake motion. Further details on the
calibration procedure of the soil dynamic
properties can be found in Pitilakis et al. 2004,
Kirtas et al. 2006 and Rovithis et al. 2006.
The coupled system response obtained by
utilizing the two different approaches of
calculating the soil springs and dashpots is
compared in Fig 9 in terms of structural and pile
response.
It is observed that when the analytical
expressions are implemented, structural
accelerations and pile bending moments seem
to be overestimated with respect to the
recorded response. Especially for the low
amplitude motion the computed structural
response resulted approximately two times
larger than the recorded one, while higher peak
bending moments were also observed.
Furthermore the superstructure responded at a
higher frequency than the actual recordings
show. On the contrary the proposed p-y model
succeeds in capturing correctly the basic
characteristics of the soil-pile system response.

Fig 8a and Fig 8b show the comparison


between the springs and dashpots obtained by
the p-y procedure and by utilizing the analytical
expressions described above. In particular, Fig
8a shows the frequency dependency of the real
part kx() of the impedance Sx normalized to
the soil modulus of elasticity Es while Fig 9b
depicts the variation with frequency of the
dashpot coefficient Cx (given in equation 5)
corresponding to the sum of material and
radiation damping. The satisfactory agreement
that is observed throughout the examined
frequency range verifies further the applicability
of the adopted procedure. Indeed the damping
term is reduced with increasing frequency while
the real part of the dynamic soil stiffness is
practically insensitive to the frequency content
of the excitation force with an exception at the
low frequency range where a slight difference is
observed. This should be correlated with
resonance effects that dominate the response
at the first natural frequency of the soil deposit.
Seismic excitation at the bedrock level

The experimental recordings from the


centrifuge test used in the validation process
were further utilized in order to investigate the
effect of the soil springs and dashpots
coefficients on the coupled soil-pile-structure
system response where both kinematic and
inertial
soil-pile-structure
interaction
is
incorporated. For this reason, the BDWF model
shown in Fig 5b was reanalyzed using spring
and dashpot coefficients calculated from the
existing analytical expressions and the
computed response was compared to the one
obtained utilizing the p-y loops for the
determination of the springs and dashpots.

C (KNsec/m)

P-Y analysis
Gazetas and Dobry (Eq.12)
Kavvadas and Gazetas (Eq.17)

2.5
=Kx/Es

Gazetas and Dobry (z>2.5D)


P-Y analysis
Gazetas and Dobry (z<2.5D)

10000

1.5
1

8000
6000
4000
2000

0.5

f (Hz)

f(Hz)

0
1

0
10

(a)

10

(b)

Fig 8: Comparison of analytical methods and p-y analysis results regarding real (a) and imaginary
(b) part of the pilehead dynamic stiffness (Homogeneous soil profile, fixed head pile, Ep/Es=1000,
L/D=20, s=0.05)

365

0.2

0.3

Acc. (g)

0.6

0
tim e (sec)

-0.4
0

10

0.8

0.6
0.4

0.05g
0.20g

8
x`

14
16
Centrifuge

1
T (sec)

12

0.2

10

1.5

0.5

1.5

10 12 14 16

BDWF model
(p-y procedure)
BDWF model
(Eq.12-16)

tim e (sec)

-0.6

12

0.5

-0.3

Centrifuge

Sa(g)

Mom ent (MNm )

z(m)

-0.2

Sa(g)

Acc. (g)

0.4

(a)

2
T (sec)

BDWF model (p-y proc.)


BDWF model (Eq.12-16)

(b)

(c)

Fig 9: Comparison of the coupled system response utilizing analytical expressions and p-y procedure to
estimate springs and dashpots a) Pilehead response spectra and superstructure acceleration time
history (0.05g) b) Pilehead response spectra and superstructure acceleration time history (0.20g) c)
Peak pile bending moments

mechanism. On the other hand, utilizing


existing analytical springs and dashpots
formulas to the examined coupled system
showed that under certain conditions (i.e
presence of sharp soil stiffness discontinuities
combined with strong inertial interaction) pile
and structural response may be overestimated
and hence further improvement of the existing
analytical expressions is needed.

Although the above remarks should not be


generalized, they indicate that a more rigorous
estimation of distributed springs and dashpots,
as it is proposed and developed herein, may be
required when the coupled soil-pile-structure
system with a particular soft foundation soil is
investigated.
CONCLUSIONS

The present paper focuses on the


evaluation of dynamic soil pile interaction
based on back calculated p-y loops,
emphasizing on the computation of frequency
dependent springs and dashpots that could be
employed within the framework of a beam on
Dynamic Winkler foundation modeling. The
adopted procedure was validated with
centrifuge tests results and was further
compared to existing analytical formulas of soil
springs and dashpots commonly utilized in
seismic soil-pile interaction studies. The
application of the procedure to different soil-pile
systems as well as to different loading
scenarios revealed that p-y loops provide a
satisfactory estimation of frequency dependent
springs and dashpots which in turn can
adequately reproduce soil-pile interaction

ACKNOWLEDGEMENTS

This work is part of the research project XSOILS (Foundation of Engineering Structures in
Seismically Problematic Soils under Strong
Ground Excitations) funded by the Hellenic
General
Secretariat
of
Research
and
Technology. The first author would also like to
acknowledge the contribution of IKY foundation.
REFERENCES
Ahmad I., El Naggar H., Khan A.N., "Fixed head
kinematic pile bending moment time history:
Artificial neural network approach" Joint
International Conference on Computing and
Decision Making in Civil and Building

366

Engineering, June 14-16, Montreal, Canada,


2006

Kirtas E., Rovithis E., Pitilakis K. and Sextos A.,


"Numerical Investigation of Potential Foundation
Intervention as a means for Mitigating Seismic
Risk", Proc. 8th U.S. National Conference on
Earthquake Engineering, April 18-22, San
Fransisco, Paper No.833, 2006

American Petroleum Institute (API). Recommended


Practice
for
Planning,
Designing
and
Constructing Fixed Offshore Platforms Working
Stress Design, API Recommended Practice
2AWSD (RP 2A-WSD), 20th edition, 191p, 1993

Lok T.M., Pestana J. and Seed R.B., "Numerical


modeling and simulation of coupled seismic soilpile-structure
interaction"
Geotechnical
Earthquake Engineering and Soil Dynamics,
Vol.III, Dakoulas P. Yegian M and Holtz R (eds),
Geotechnical Special Publication, No.75, ASCE,
pp. 1211-1222, 1998

Angelides D.C., and Roesset J.M., "Nonlinear lateral


dynamic stiffness of piles" Journal of the
Geotechnical Engineering Division, ASCE,
Vol.107, No. GT11 (1981)
ANSYS Inc (2000) ANSYS Users manual. Version
8.1, SAS IP, Houston, USA.

Makris N., Badoni D., Delis E., Gazetas G.,


"Prediction of observed bridge response with
soil-pile-structure
interaction"
Journal
of
Structural Engineering, 120(10), 2992-3011,
1994

Badoni D. & Makris N., "Nonlinear response of


single piles under lateral inertial and seismic
loads," Soil Dynamics and Earthquake
Engineering, 15, 29-43, 1995

Makris N., Tazoh T., Yun X. & Fill A.C., "Prediction


of the measured response of a scaled soil-pilesuperstructure system" Soil Dynamics and
Earthquake Engineering, 16, 113-124, 1996

Boulanger R.W., Curras C.J., Kutter B.L., Wilson


D.W., and Abghari A., "Seismic soil-pile-structure
interaction experiments and analyses," Journal of
Geotechnical
and
Geoenvironmental
Engineering, 125(9),750-759,1999.
Dobry R., Vicente E., O'Rourke E.. and Roesset J.,
"Horizontal stiffness and damping of single piles",
Journal of the Geotechnical Engineering Division,
ASCE, Vol.108, No.GT3, 1982
Finn W.DL, Barton Y.O., and Towhata I., "Seismic
response of offshore pile foundations," Proc. 7th
Pan-American Conference on Soil Mechanics
and Foundation Engineering, Vancouver, B.C,
1983

Makris N. and Gazetas G., "Dynamic pile-soil-pile


interaction. Part II: Lateral and seismic response"
Earthquake
Engineering
and
Structural
Dynamics, Vol.21, 145-162, 1992
Matlock H., "Correlations for design of laterally
loaded piles in soft clay," Proc, 2nd Annual
Offshore Technology Conference, Vol.1, 577-594,
1970
Meymand P.J., "Shaking table scale model tests of
nonlinear soil-pile-superstructure interaction in
soft clay" PhD Thesis, University of California,
Berkeley, 1998

Gazetas G. and Dobry R., "Horizontal response of


piles in layered soils," Journal of Geotechnical
Engineering, ASCE, 110(1), 1984

Murchinson J.M. and O'Neill M.W., "Evaluation of py relationships in cohesionless soils," Analysis
and design of pile foundations, ASCE, J.R Meyer,
ed., 174-191, 1984

Gazetas G., Fan K., Tazoh T., Shimizu K., Kavvadas


M. and Makris N., "Seismic pile-group-structure
interaction" Piles under dynamic loads,
Geotechnical
Special
Publication
No.34,
S.Prakash (ed), ASCE, 1992

Mylonakis G., Nikolaou S. and Gazetas G., "Soilpile-Bridge seismic interaction: Kinematic and
inertial effects. Part I: Soft soil" Earthquake
Engineering and structural Dynamics, 26, 337359, 1997

Georgiadis M., Anagnostopoulos C. and Saflekou S.,


"Centrifugal testing of laterally loaded piles in
sand," Canadian Geotechnical Journal, 29, 208216, 1992

Mylonakis G., "Contributions to static and seismic


analysis of piles and pile-supported bridge piers"
PhD Thesis, State University of New York,
Buffalo, 1995

Janoyan K., Stewart J.P. and Wallace J.W.,


"Analysis of p-y curves from lateral load test of
large diameter Drilled Shaft in Stiff Clay,"
Proceedings 6th Caltrans Workshop on Seismic
Research, Sacramento, CA, 2001

Pitilakis K., Kirtas E., Sextos A., Bolton M.,


Madabhushi G., Brennan A., "Validation by
centrifuge testing of numerical simulations for
soil-foundation-structure systems", Proc.13th
World Conference on Earthquake Engineering ,
Vancouver, Canada, 2004

Kavvadas M. and Gazetas G., "Kinematic seismic


response and bending of free-head piles in
layered soil," Geotechnique, 43(2), 207-222,
1993

367

Reese L. C., and Welch R. C., (1975). Lateral


loading of deep foundation in stiff clay. Journal
of Geotechnical Engineering Div. ASCE, 101(7),
pp 633-649.
Reese L.C., Cox W. R. and Koop F. D., (1974).
Analysis of laterally loaded piles in sand Proc.,
6th Offshore Technology Conf., Paper 2080,
Houston, Texas, 473-483.
Rovithis E., Pitilakis K., Kirtas E., "Seismic response
of piled foundations in soft soil formations" Proc.
5th Hellenic conference on Geotechnical and
Geoenviromental Engineering, Vol. II, 433-440,
2006 (In Greek)
Rovithis E., Kirtas E. and Pitilakis K., (2007a)
"Evaluation of Dynamic soil-pile interaction
based on back-calculated P-Y curves" 4th
International
Conference
on
Earthquake
Geotechnical Engineering, Thessaloniki, Greece,
(submitted)
Rovithis E., Kirtas E. and Pitilakis K., (2007b)
"Insight
into
soil-pile-structure
interaction
mechanism including inertial and kinematic
effects" 4th International Conference on
Earthquake
Geotechnical
Engineering,
Thessaloniki, Greece, (submitted)
Ting J.M., "Full-scale cyclic dynamic lateral pile
response," ASCE Journal of Geotechnical
Engineering, 113(1), 30-45, 1987
Wilson DW, Boulanger RW, and Kutter BL. "Soilpile-superstructure interaction at soft or
liquefiable soil sites - Centrifuge data report for
Csp4," Rep. No UCD/CGMDR-97/05, Center for
Geotechnical Modeling, Department of Civil and
Environmental
Engineering,
University
of
California, Davis, California, 1997
Wilson D. W., "Soil-pile-superstructure interaction in
liquefying sand and soft clay, PhD Thesis,
Department of Civil and Environmental
Engineering, University of California, Davis, 1998

368

Linear elastic transient response of deep rigid foundations:


From 3D finite element simulations to design
Varun1, D. Assimaki2 & G. Gazetas3
1

Graduate Research Assistant, Georgia Institute of Technology, School of CEE, Atlanta GA


2
Assistant Professor, Georgia Institute of Technology, School of CEE, Atlanta GA
3
Professor, National Technical University of Athens, GREECE

Abstract
The transient response of drilled shafts or pier foundations, namely large rigid foundation
elements of length-to-diameter aspect ratio D/B=2-6, is characterized by a complex stress
distribution at the pier-soil interface that cannot be adequately represented by means of
existing models for shallow foundations or flexible piles. On the other hand, while 3D
numerical solutions are feasible, they are infrequently employed in practice due to their
associated cost and effort. Prompted by the lack of simplified models for the design of
drilled shafts in current practice, we here develop an analytical model that accounts for the
multitude of soil resistance mechanisms mobilized at their base and circumference, while
retaining the advantages of simplified methodologies for the design of non-critical facilities.
The characteristics of soil resistance mechanisms and corresponding complex spring
functions are developed on the basis of finite element simulations, by equating the stiffness
matrix terms and/or overall numerically computed response to the analytical expressions
derived for the proposed Winkler model. Numerical simulations of the transient system
response to vertically propagating shear waves are successively compared and found to be
in agreement with the analytically predicted response. The formulation of frequency-dependent complex spring functions including material hysteretic energy absorption is also described, while extension of the methodology that accounts for nonlinear soil behavior and soilfoundation interface separation is described in the conclusion and is being currently
undertaken by the authors.

storey parking decks and scour vulnerable


structures (ONeil and Reese, 1999).
Caissons are highly versatile in constructability for a wide variety of soil formations,
and can be installed in virtually any soil type
including residual soils, karstic formations,
soft soils and marine sites. Even further, since
no dewatering is necessary during construction, caissons are particularly advantageous
at soft sites or sites where excessive groundwater is considered to be critical for the selection of the excavation and support method.
The high capacity of single caisson elements in axial as well as lateral loading, and
the ability to connect directly to structural
members without caps enables them to effectively replace pile groups, and makes them a
popular choice for structures anticipated to be
subjected to significant lateral loads. They are

INTRODUCTION

Drilled shafts or pier foundations are large


foundation blocks of intermediate length-todiameter aspect ratio (typically within the
range D/B=2-6), whose diameter ranges from
2 to 12 feet (ONeil and Reese, 1999). Prefabricated and sunk in place, or cast in-situ,
large diameter caisson foundations are
typically used as bridge foundation elements,
deep-water wharves, and overpasses, and
among others, examples include the Rokko
Island Bridge in Kobe, Japan and the
Brooklyn Bridge in New York, NY. Small
diameter caissons on the other hand are
extensively encountered either as single
foundation components of transmission towers and heliostats, or in groups as part of the
foundation system of high rise buildings, multi-

369

simulating the overall stiffness of the soil-foundation system. While the first condition resembles the approach followed within the context
of shallow foundation theories, the latter
renders p-y curve approaches to be more
suitable when accounting for layered media.
Results presented in this paper show that
caisson foundations are indeed expected to
behave as rigid elements similarly to shallow
foundations for maximum depth to diameter
ratios of 6 and typical soil-caisson impedance
contrasts, while beyond that range of aspect
ratios, their response resembles that of
flexible piles. Nonetheless, the embedded
foundation solutions are shown to be
applicable only for low embedment ratios (D/B
< 2).
For pier foundations with intermediate
length (D/B=2 to 6), the soil-structure interaction effects comprising the load-transfer
mechanisms from the superstructure to the
surrounding soil and the potential altering of
loads transferred through the foundation from
the soil to the structural elements (e.g. during
seismic motion) are associated with a
complex stress distribution at the pier-soil
interface with comparable contributions from
the base and the shaft that cannot be
captured by shallow of deep foundation simplified approaches.
Prompted by the lack of a simplified
design methodology for caisson foundations
that may be used to adequately predict their
dynamic response at intermediate levels of
target design sophistication, we here develop
a dynamic Winkler model that properly
accounts for the multitude of soil resistance
mechanisms mobilized at the base and the
circumference of laterally loaded piers -thus
retaining the advantages of Winkler-type models will allow for realistic representation of
the complex soil-structure interaction effects
associated with these foundation elements.

thus used worldwide by agencies that focus


on the design and construction of lifelines in a
wide variety of site conditions, such as the
Departments of Transportation.
Categorized according to their geometry
characteristics as intermediate embedded
foundations when compared to shallow and
deep
elements
(Figure
1),
caisson
foundations are currently designed by means
of: (i) existing shallow embedded foundation
methods; typical examples include the analytical, semi-analytical and numerical approaches by Novak & Beredugo (1972), Kausel &
Roesset (1975), Elsabee & Morray (1977),
Dominguez (1978), Tassoulas (1981), Mita &
Luco (1989), Tajirian & Tabatabaie (1985) and
Gazetas (1983), and co-workers (2006), the
majority of which have been developed for
cylindrical foundations (note that the latter
may be applied to foundation elements of
arbitrary cross-sections); or (ii) flexible pile
approaches (also referred to as p-y and t-z
curves) developed semi-empirically as a function of the soil type, e.g. Lam and Chaudhury
(1997). Alternatively, while three-dimensional
numerical solutions are feasible, their
application for the design of non-critical facilities is typically prohibited by the associated
site investigation cost, computational time,
and user expertise required.

Fig. 1 Foundation categorization as a function of


the geometry characteristics

The comparable dimensions of depth to


diameter of caisson foundations imply that
within the context of assessing the global
foundation stiffness, neither the circumference
nor the base resistance mechanisms may be
neglected. On the other hand, however, pier
foundations typically extend through layered
soil formations, a fact that requires analytical
solutions to be capable of capturing the
vertical variability in soil stiffness when

SIMPLIFIED MODEL DESCRIPTION

Figure 2 schematically depicts the stress


distribution and associated stress resultants
developed at the foundation-soil interface,
when a typical caisson is subjected to
transverse loading at the top, the former here
represented by a combination of a lateral
concentrated load (V) and a moment (M). As

370

element simulations, comprise the following:


(i) lateral translational springs representing
the lateral force-displacement soil response
(kx); (ii) rotational springs representing the
moment developed at the centerline of pier
due to vertical shear stress acting at the
perimeter of pier, induced by pier rotation (k);
(iii) base translational concentrated spring
representing the horizontal shear force-base
displacement response (kbx); and (iv) base
rotational spring representing the moment due
to normal stress acting at the base of pier,
induced by base rotation (kb).

can be readily seen, four mechanisms are


identified as significantly contributing to the
pier response (Mayne et al., 1992). The
mathematical expressions for the resistance
mobilized by these mechanisms are
presented below and comprise the following:
(a) Lateral resistance per unit length due to
normal stresses along the shaft:
2

Ph =

cos + r sin rd

(b) Resisting moment per unit length due to


vertical shear stress along the shaft:
2

Mh =

rz

(B / 2)2 cos d

(c) Lateral base resistance due to horizontal


shear stress:
Pb =

B
2 2

rz

cos + z sin )rd dr

0 0

(d) Base resisting moment due to normal


stresses:
Mb =

B
2 2

cos )r 2 d dr

0 0

rz
y
Lateral
resistance
below zr

x
Vertical
side
resistance

The model is based on the assumption


that the response of each soil layer is
decoupled from the overlying and underlying
ones, namely the shear beam deformation
simplification proposed by Novak et al. (1978).
As a result, in absence of coupling between
adjacent soil resistance mechanisms, the total
response can be obtained through integration
of the total resistance offered by the individual
springs for each layer. While this assumption
is not valid in the immediate vicinity of soil
layer interfaces, for soil layers where the
cross-layer interaction region represents a
small percentage of the total layer thickness
(i.e. adequately thick layers), the coupling
effect diminishes very rapidly with distance
from the interface and the overall contribution
of the coupling to total response becomes
practically negligible.
In the ensuing, the deformed configuration
of the model shown in Figure 3 subjected to
transverse loading is investigated by means of

Lateral resistance
above zr

Point of rotation zr

Lateral flow

rz

Fig. 3 Four spring model implemented in this study


for the macroscopic representation of the
complex resistance mechanisms developed
at the soil-caisson interface (modified from
Assimaki et al, 2001)

Tip shear
resistance

Tip axial
resistance

Fig. 2 Primary resistance mechanisms contributing


to the overall stiffness of caisson foundations subjected to transverse loading at
ground surface; the angle above represents the polar coordinate variable of integration (modified from Mayne et al, 1992)

Following Mayne et al. (1992), Assimaki et al.


(2001) and Gerolymos & Gazetas (2006), a
four spring model is here implemented to capture macroscopically the aforementioned resistance mechanisms (Figure 3). The distributed and concentrated spring functions, calibrated in the ensuing by means of 3D finite

371

the Euler beam theory formulation, and simplified deformation profiles are developed for typical geometry characteristics and foundationsoil impedance contrasts. The overall foundation stiffness matrix as interpreted from the
top of the caisson by integration of the individual resistance mechanisms is successively
evaluated based on the deformation simplifications.

distributed spring constants are proportional


to the soil stiffness and foundation geometry
characteristics as k x Es and k Es B 2 . As a
result, for typical caisson geometries and soilfoundation impedance contrasts:
k

E pIp

= a ib = (c id )

Considering that in the most general case, the


response of a caisson subjected to transverse
loading at the top can be approximated by the
Euler beam theory, the governing equation for
static deformation of a caisson may be formulated as follows:
z

a=

,c =

E p I p u (0) = M
E p I p u (0) = ( V ) = V

(3)

E p I p u (D ) = ( kb u '(D ) )
E p I p u (D ) = ( kbx u(D ) )

(4)

(9)

(10)

Note that neglecting the contribution of the


base as part of the resistance mechanisms of
the overall foundation stiffness, namely defining:
u ( z = D ) 0, u ( z = D ) 0 ,
the deformed configuration solution is simplified to the well-known long pile solution as follows:

The general solution constants are given by


the solution of the biquadratic equation (5) as
follows:

u( z ) = A1e cz cos(dz ) + A2 e cz sin(dz )

k
x
E
pIp

(8)

subjected to the following boundary conditions


at the top (z = 0) given by Eq.9, and base (z =
D) given by Eq.10 of the caisson:

The general solution of the 4th order differential equation above is u( z ) = e z , which substituted in (4) results in the following expression:
E p I p 4 + k 2 + k x = 0
(5)

1 k
1 k

2 EpIp
4 E p I p

+ A4 e c ( D z ) sin(d (D z ))

For the case of a uniform cylinder, Eq.2 and 3


result in the following:

(a 2 + b 2 ) a
2

+ A3 e c ( D z ) cos(d (D z )) +

where M and V are the concentrated moment


and lateral force at the top of the foundation, u
is the lateral deformation variable of the beam
with depth, z is the depth from ground
surface, and Ep and Ip are the foundation
Youngs modulus and area moment of inertia
correspondingly. Differentiating Eq.1 with respect to z, we get:

E p I p u IV + k u + k x u = 0

(a 2 + b 2 ) + a
,d =
2

u( z ) = A1e cz cos(dz ) + A2 e cz sin(dz ) +

(2)

2
*
( EpIpu ) = zM2

k
1 k
1 k
,b = x
2 EpIp
E p I p 4 E p I p

The general solution of the beam in the deformed configuration can be written as:

(1)

M * = M + Vz k u I ( )d k x u( )( z )d

(7)

Based on the simplification introduced by


Eq.7, Eq.6 may be further simplified and the
resulting solution constants may now be formulated as:

APPROXIMATE CAISSON DEFORMATION


BASED ON EULER BEAM THEORY

E p I p u = M * ,

k
<< x .
EpIp

(11)

and is subjected to the boundary conditions at


the top as described by Eq.9. In this case, the
resistance mechanisms at the pile-soil interface may be described by means of the distributed translational spring kx along the pile

(6)

Based on previous studies by Assimaki et al


(2001) and Gerolymos & Gazetas (2006), the

372

the base has a significant effect on the


response and hence cannot be neglected;
(ii) The range of D/B ratio for which the
response may be approximated by a rigid
body is also a function of the impedance
contrast Ep/Es; note that for typical reinforced
concrete or steel casing, this ratio is on the
order of 104-105; and
(iii) For stiffness ratios on the order of 104-105,
the variation in rotation along the length is
less than 5% for aspect ratios D/B<6, rendering the rigid body approximation valid.

k
= x
4E I
pp

(12)

(13)

Employing series expansion of the products:


e z cos( z ) = 1 2 z + O2 power terms
e z sin( z ) = z + O2 power terms

the following further simplifications of the solution may be introduced based on the dimensional analysis described above ( k x Es and
k Es B 2 ):
Ep
kx

Ep
Es

>>;

D
1

Ip 4

(14)
0.4
0.6

4
D << 1

D/B=2
D/B=4
D/B=6
D/B=8
D/B=10

0.2

and as a result, the maximum value that can


be attained by the exponent term (z) is:
k
max { z} = x
4E I
pp

Aspect ratio (D/B), Ep/Es=104, V=1, M=0


0
0.2
0.4
0.6

0.8

0.8

(15)

u/umax

1.0
-0.5

Finally, introducing Eq.15 into Eq.12 and neglecting the higher order terms, the deformed
configuration conditioned on the approximations described by Eq.14 may be described by
a rigid body translation and rotation, namely:
u ( z ) = ut z
(16)

0.5

u/umax

1.0
1.0
0.50

0.75

1.00

Stiffness Contrast (Ep/Es), D/B=6, V=1, M=0


0

0
3

Ep/Es=10
4
Ep/Es=10
5
Ep/Es=10

0.2
0.4

where ut is the translation at the top of the


foundation and is the rigid body rotation. As
can be readily seen, for high pier-soil impedance contrast (Ep>>Es) and low aspect ratios
(D/B) the response may be approximated by
the rigid body motion described in Eq.16.
Implementing in Eq.8 (4-spring model) and
Eq.11 (2-spring model) characteristic values
for the distributed and concentrated springs,
selected here as the static plane strain solution proposed by Novak et al. (1978) for the
former and the surface circular foundation
stiffness on elastic halfspace proposed by
Veletsos & Wei (1971) and Veletsos & Verbic
(1974) for the latter, Figure 4 depicts the
predicted response of the model as a function
of the aspect ratio (D/B) and the impedance
contrast (Ep/Es). Based on the results shown,
it can be readily seen that:
(i) While for large aspect ratios (D/B>8) the 4spring response may be approximated by the
pile solution described by Eq.11, for D/B<8

z/D

+ M cos( z ) M sin( z ) ,

z/D

e
2 2E p I p

0.2
0.4

z/D

u=

z/D

length, yielding the following solution:

0.6

0.6

0.8

0.8

1.0
-0.2

u
0

0.2

0.4

1.0
0.6 -0.20 -0.15 -0.10 -0.05

Fig. 4 Normalized deformation and rotation of


beam on linear elastic 4-spring Winkler
soil as a function of: (top) the aspect ratio
(D/B), and (bottom) and the pier-soil stiffness contrast (Ep/Es), for a unit lateral load
(V=1, M=0) applied at the top.
STIFFNESS MATRIX FORMULATION:
RIGID BODY DEFORMED APPROXIMATION

As shown above, the caisson is anticipated to


respond as a rigid body for typical ranges of
soil and foundation stiffness, and aspect ratios
D/B < 6. Therefore, upon the application of a
lateral force (V) and an overturning moment
(M) at the top, the net translation and rotation

373

exerted by the caisson are depicted by Figure


5, and may be adequately described in terms
of the displacement at the top (ut) and the
rotation angle of the rigid pier ().
In the ensuing, the dynamic springs are
formulated in a complex form representing the
static stiffness potentially altered when the
foundation is subjected to dynamic loading,
and the radiation damping of energy away
from the foundation during cyclic vibrations:
K * = K stat k '(ao ) + iaoC(ao )

and evaluating moment equilibrium at the top


of the caisson:
D

*
M * = k x* u * ( z )zdz kbx
u * (D ).D
0

D
*
k bx
M * = ut* k x* 2 m
D
2

(17)

D3
*
+ * k x* 2 m
+ k bx
D 2 + k* D + k b*
3

where Kstat is the static (push-over) stiffness


ao = B / Vs is the dimensionless frequency with
, B, and Vs being the loading frequency, pier
diameter and soil shear wave velocity
correspondingly, k (ao ) is a frequency dependent stiffness coefficient, and C(ao ) is the
frequency dependent radiation damping component of the complex stiffness.

*
V * K xx
=
* *
M K rx

K xr* ut*

K rr* *

K xr* = K rx* = k x* 2 m

K rr* = k x* 2 m

The dynamic response at any point along the


caisson is approximated by Eq.16, namely:
u * ( z ) = ut* * z ,
(18)
where (*) indicates the complex response, namely the amplitude and phase of the top displacement and rotation. Successively, requiring dynamic equilibrium of forces in the horizontal direction:
0

*
V * = k x* u * ( z )dz + kbx
u * (D ) + mu* ( z )dz

D
3

*
V * / EB 2 K xx
/ EB
=

*
2
3
M / EB K rx / EB

*
+
V * = ut* k x* 2 m D + k bx

D2
*
+ * k x* 2 m
kbx
D
2

D2
2

*
+ k bx
D

*
+ kbx
D 2 + k* D + kb*

(23)
(24)

As can be readily seen from the expressions


above, the assumption of a rigid caisson response results in a symmetric stiffness matrix
for the overall foundation system.
Following the dimensional analysis described by Assimaki et al (2001) and Gerolymos
& Gazetas (2006), the complex stiffness terms
in Eq.21 are normalized with respect to the
Youngs modulus of soil (Es) and the diameter
of pier (B), namely the stiffness and geometry
characteristics of the surrounding soil and
foundation element respectively. In a dimensionless form, the equations are given as
follows:

Fig. 5 Rigid body response of pier upon application of lateral load and overturning moment

(21)

where the left-hand side of the equation


represents the dynamic forcing function (F*)
applied at the top of the foundation, and the
right-hand side corresponds to the product of
the complex stiffness matrix (K*) of the soilfoundation system as interpreted from the
foundation top to the response vector (U*),
namely the displacement and rotation of the
caisson. In Eq.21, the individual components
of the stiffness matrix correspond to the
following expressions:
*
*
K xx
= ( k x* 2 m ) D + kbx
(22)

(20)

the following expression is formulated for the


dynamic stiffness matrix, when expressing
Eq.19 and Eq.20 in a matrix form:

M
V

ut

+ k* * dz + k b* * mu* ( z )zdz

(19)

where:

374

K xr* / EB 2 ut* / B

K rr* / EB 3 *

(25)

*
*

K xx
k x* D kbx

= +
EB E B EB

8(1 + )
s

1 k x*
K xr*
=

2
EB
2 E

formation and pier are simulated using 3D


continuum soil elements (8 node brick
elements), while for the purpose of this work,
linear
elastic
material
models
were
implemented and perfect bonding was
assumed at the interface, i.e., no separation
was allowed under tensile stress. This
assumption is equivalent to that of a
cylindrical foundation welded in soil, as
described by Kausel (1974). Mesh sensitivity
studies were conducted for the element size,
far-field shape and far-field distance. For the
static simulations, comparison between DYNAFLOW simulations with fixed far-field and ABAQUS simulations with infinite elements in the
far-field indicated that implementation of
lateral far-field distance of 5B is not sufficient
for the representation of the problem. Results
indicated that increasing the far-field distance
to 10B provided sufficiently accurate results
with less than 5% discrepancy in the estimated response by means of the two alternative solutions. Also, two types of meshes were
compared (Figure 6), Mesh A with cylindrical
far field and adaptive element size and mesh
B with rectangular far-field and constant
element size; while the difference in results
was negligible conditioned on the adequate
distance of the far-field, the adaptive mesh A
was selected due to the implied increased
computational efficiency.

(26)

2 D
ao
B

*
D
D k bx

+
B EB B
2

2
p 2 D
1
ao
+

2 8(1 + ) s B

1 k * D 3 k * D 2
x + bx
K rr* 3 E B EB B
=

3
EB
k* D k*
... + 2 + b3

EB B EB

(27)

(28)

3
p 2 D
1

ao
3 8(1 + ) s B

3D FINITE ELEMENT FORMULATION


The stiffness matrix described by Eq.25, can
be evaluated numerically by computing the
displacements at the top and rotations of the
pier resulting from the application of a unit lateral force and a unit overturning moment, and
inverting the response matrix as shown below:
ut K xx
= K
rx

K xr V uV
=
K rr M V

uM V
M M

(29)

This method, referred to as the flexibility


approach and extensively applied in the field
of structural mechanics, will be employed in
the ensuing for the calibration of the distributed and concentrated springs applicable for
caisson foundations in the aspect ratio range
D/B<6 by means of 3D finite element simulations, in absence of closed form solutions for
the overall stiffness matrix to be used for this
purpose.
The simulations for this study are
performed using the finite element software
package DYNAFLOW (Prvost, 1983) and verified, for the case of static loading using the
software package ABAQUS (Hibbit et al,
2006). Taking advantage of the symmetry of
the problem, only half of the numerical domain
is simulated, thus reducing considerably the
required computational effort. Both the soil

Fig. 6 Sensitivity analysis of finite element mesh


discretization and location of far-field boundaries: Mesh A with cylindrical far field and
Mesh B with rectangular far-field

375

Nonetheless, implementation of the fixed


boundary conditions in numerical simulations
of dynamic or transient response problems
would result in spurious reflections of the elastic waves back into the truncated numerical
domain, an effect particularly pronounced in
the foregoing simulations due to the absence
of material damping. On the other hand, the
use of far-field viscoelastic dampers or infinite
elements was shown to yield unsatisfactory
results due to the large angle of incidence of
waves at the boundaries for the 3D finite
element problem under investigation.
For this purpose, a new type of boundary was
implemented in this study, hereby referred to
as sponge boundary and schematically
depicted in Figure 7. In particular, the reflection of outgoing waves back into the region
of interest is avoided by enclosing the region
in the sponge layer with progressively increasing damping coefficients. The mechanical
sponge layer-soil impedance is approximately
equal to unity thus minimizing the generation
of reflected waves at the interface, while
energy absorption is represented by means of
Rayleigh damping increasing with distance to
avoid any spurious reflections due to sudden
change in impedance.

accuracy is described by the following


expression:
B 2 fB 2 B 2
ao =
=
=

(30)

vs
vs
6
For the case of Model A, the ratio of the
maximum element size to foundation width is
=0.67, and therefore the maximum dimensionless frequency represented is approximately amax3.5. Based on the aforementioned
highest accurately represented frequency, the
minimum time step is consequently given by
the following expression:
B B 1 Tmin
=

=
(31)
t=
6
vs
f 6f
In the simulations presented in the ensuing of
this study, a time step of Tmin/20 to Tmin/40 has
been employed thus ensuring the accurate
representation of the propagating wavelengths.
Comparison with published formulations
It has been shown that based on the assumption of rigid body motion of the pier and consequent absence of flexural bending representation, the stiffness matrix is symmetric,
namely the off-diagonal coupled stiffness
terms are equal Kxr=Krx. Figure 8 depicts the
sensitivity of the predicted system response to
the coupled stiffness Kxr=Krx, namely the
variation of the expression:
dK xr
(K K rx )
= 2 xr
(K xr + K rx )
K xravg

(32)

as a function of the aspect ratio D/B. for soil


Poisson ratio =0.3 and foundation-soil impedance ratios Ep/Es=104 and 105, representative of the cases of concrete and steel pier
respectively.
Fig. 7 Finite element mesh of the numerical model,
depicting the truncated computational domain and energy absorbing (sponge) layer.

The numerically accurate representation of


wave propagation problems susceptible to numerical attenuation because of under sampling requires at least 6-7 elements per wavelength. Based on this requirement, / 6 B
where = ratio of size of largest element to B.
As a result, the maximum dimensionless
frequency that can be simulated with sufficient

10

-1

10

-2

10

-3

10

-4

Ep/Es=10

Ep/Es=10

10

-5

10

Fig. 8 Deviation of pier from rigid behavior for


different Ep/Es: coupled stiffness sensitivity
as a function of the aspect ratio D/B for
two foundation-soil impedance contrasts.

376

Note that for D/B ratios larger than 6, the


deviation is shown to exceed 0.05 (5%) for
Ep/Es=104, and as a result, the assumption of
a rigid pier is considered valid only for aspect
ratios D/B 6.
Figures 9a-c show the comparison of
results obtained for the global stiffness matrix
of the caisson foundation via 3D finite elements with existing formulations by Kausel
(1974), Wolf (1997) and Elsabee & Morray
(1977) developed for shallow foundations, by
Davidson (1982) and Gerolymos & Gazetas
(2006) for caisson foundations, and by means
of a four spring model with base springs equal
to the expressions by Veletsos and Verbic
(1974) for a surface foundation on elastic
halfspace, and shaft resistance by Novak et al
(1978). As can be readily seen, the models
proposed by Kausel (1974), Elsabee & Morray
(1977), Wolf (1997) and Davidson (1982) fail
to capture at least one of the three global stiffness terms for the foundation system. On the
other hand, stiffness values predicted by
springs obtained from combination of formulations by Novak et al (1978) and Veletsos
and Verbic (1974), namely constant spring
values independent of the foundation aspect
ratio (D/B) based on plane-strain theory, is a
good approximation to the configuration
investigated here. Nonetheless, while the
latter combined approach captures qualitatively the overall variation of stiffness as a
function of D/B, it quantitatively predicts lower
stiffness values than the actual ones obtained
by means of 3D finite element simulations.
Figure 9d shows the location of center of
rotation of pier as a function of eccentricity of
applied loading obtained from 3D simulations
and by means of the Euler beam theory (4spring formulation). It should be also noted
herein that results showed that the 4 spring
approximation of the pier may capture the true
response much better than the 2 spring
model, while the observed deviation from the
2-spring model was shown to increase as D/B
ratio decreased and base effects would become more pronounced.
Based on the aforementioned conclusions,
it is clear that none of the existing models may
be used to capture all the three modes of soil
resistance completely. There exists therefore
a clear need to calibrate the springs of the
proposed model, which may be successively

employed for the evaluation of the overall


foundation stiffness; evaluating the lateral,
rocking and coupled stiffness at the top of the
foundation; which may successively be used
in analyses of structural response to replace
the continuum formulation of the infinite
domain and foundation element by the
foundation-soil stiffness matrix.
Kxx/EB

20
15
10

5
D/B

Kxr /EB^2

0
80
60
40
20
D/B
0

Krr/EB^3

500
400

Novak (1967) Velestos(1971)


Elsabee-Kausel(1977)
Wolf (1997)
EPRI (1982)
Gerolymos-Gazetas(2006)
3D simulations

300

200
100

D/B
0
0
1.1

2
2 (4s)
4 (4s)
6 (4s)

2 (3D)
4 (3D)
6 (3D)

z c/D

0.9

0.8
0.7
0.6

VD/M
0.5
0

10

Fig. 9 Comparison of global stiffness matrix components of foundation as interpreted from


the top of the caisson between FE simulations and embedded foundation published
formulations; Location of center of rotation

377

from 3D FE analysis and analytical solution


(4-spring beam theory)

increases, the contribution of k and kb to the


global rotational stiffness Krr compared to kx
and kbx decreases. As a result, Krr becomes
increasingly insensitive to changes in values
of these two springs for higher D/B ratios, a
fact that renders the interpretation of their
corresponding functions by means equation
(28) (i.e. equating the analytically and numerically derived global rotational stiffness (Krr)
to derive the variation of kb and k) cumbersome for large values of the aspect ratio. As
shown in Figure 10, these values correspond
to D/B=0.75 for kb, and D/B=5-6 for k.
As can be readily seen, for the aspect ratio
region beyond which the rocking stiffness becomes insensitive to changes in value of the
aforementioned springs, the corresponding
mechanism of soil resistance may be neglected altogether, which results in a simplified
analytical model of the caisson foundation.
Based on this interpretation, the response of
pier can be broadly classified into three main
zones depicted in Figure 10: Zone I for D/B =
0-2 (four spring model), Zone II for D/B=2-6
(three spring model) and Zone III for D/B>6
(two spring model). The behavior of individual
springs as a function of the aspect ratio is
summarized below:
(i) The distributed lateral spring kx decreases
rapidly with D/B ratio in Zone I, beyond which
it becomes constant with a normalized value
of kx/E=1.48, comparable to the value of
kx/E=1.15 predicted by Novak et al (1978) for
very low frequencies under the assumption of
plane strain response. This behavior is most
possibly attributed to the fact that at higher
embedment depths, the soil layer response
becomes increasingly uncoupled from the
adjacent layers, and the resulting minimal
mobilization of shear resistance does indeed
approximate plane strain conditions. On the
contrary, the interaction between adjacent
layers is significant for shallow foundations
represented in Zone I, resulting in higher
resistance mobilized in between due to shear
interaction and increasing value of kx with
decreasing aspect ratio (D/B).
(ii) The base lateral spring kbx/EB evaluated
by means of finite elements is found to be
0.92 for surface foundations (D/B=0), which is
in excellent agreement with the value predicted by the formulation of Veletsos and Verbic
(1974) for the horizontal impedance of circular

Calibration of static springs


The distributed lateral springs along the length
and concentrated spring at the base of the
pier, namely kx and kbx correspondingly, can
be back calculated by equating the overall
lateral and coupled stiffness of the pier shown
in Eq. 22-23 to the numerically evaluated stiffness as:
K D+K
K D + 2K xr
k x = 2 xx 2 xr , k bx = xx

D
D

(33)

Successively, the overall rotational stiffness of


the foundation as interpreted from the top,
described in Equation (14), can be written as:
K rr (k x

D3
3

+ kbx D 2 ) = k D + kb

(34)

In order to evaluate k and kb, we evaluate


the equality described by Eq.34 for two different values of D in the immediate vicinity of
each other. The inherent assumption of this
approach is that the variation of these two
springs with (D/B) is small. The variation of all
four springs as a function of the foundation
aspect ratio is shown in Figure 10.
2.5
2.0
1.5
1.0

kx
k
Kbx
Kb

0.5
0
0

Fig. 10 Variation of distributed and concentrated


spring functions with aspect ratio D/B.

Resolving the spring functions by means


of the flexibility approach based on the finite
element-based estimated response, it is observed that the base rotation spring kb
contribution decreases significantly with D/B,
and becomes negligible for D/B> 0.75. Similar
behavior is observed for the variation of the
distributed rotational spring k for aspect
ratios D/B>5-6. The trend observed above is
interpreted to be directly stemming from the
procedure employed here for the derivation of
the spring functions: as the aspect ratio D/B

378

trated lateral and distributed rotational stiffness respectively.

foundations on elastic halfspace; for D/B>0


ratios, kbx increases, a behavior attributed to
the so-called trench effect according to which,
the soil at deeper layers is more constrained
as compared to the surface and therefore
mobilizes a higher shear resistance.
(iii) The value of the distributed rotational
spring k increases with D/B and becomes
approximately constant for 4<D/B<6. The
increase is explained by the increase in
confinement due to trench effect and higher
shear resistance mobilized at the sides. The
contribution of this mechanism of soil resistance significantly reduces for D/B >6, namely in
Zone III where the response of the foundation
element approximates the pile solution.
(iv) The base rotational spring kb /EB3 has an
initial value of 0.18 for D/B<<, which is in very
good agreement with the theoretical rocking
stiffness predicted by the formulation by Veletsos and Verbic (1974) for surface foundations. Nonetheless, since the relative contribution of this spring to total rocking stiffness is
very small, it may be neglected with no loss of
accuracy in the solution for D/B>1.
For aspect ratios D/B>6 and relative foundation/soil stiffness corresponding to concrete
piers, the response of the foundation is shown
to deviate significantly from the perfectly rigid
assumption. In this aspect ratio range, the
caisson behaves as a flexible foundation, and
the response can be estimated by means of
the p-y spring approach. Note that for a linear
elastic medium, the p-y curve is represented
by the distributed lateral spring kx.
Using the spring values of the simplified
three-mechanism analytical model obtained
by means of finite element simulations, simplified expressions are derived by means of
least-square curve fitting for the aspect ratio
range under investigation (namely D/B=2 to 6)
as follows:

2.00
1.50
1.00
0.50
0.00
2

1.50

1.00
k bx
D
0.13 + 0.67
EB
B

0.50

y = 0.1291x + 0.6687
2

= 0.9826
R 2 =R0.9826

0.00
2

2.50
2.00
1.50

k
D
0.23 + 1.11
EB 2
B

1.00
0.50

y = 0.2274x + 1.1059
2

= 0.9745
R 2 = R0.9745

0.00
2

Fig. 11 Least-square curve fitting for distributed


translational (top), base translational (middle) and distributed rotational (bottom)
spring variation as a function of D/B evaluated by means of 3D FE simulations.

Sensitivity analyses investigating the effects of Poisson ratio variability in the predicted response indicated that the overall stiffness terms are for the most part insensitive to
changes in Poisson ratio. Therefore, the variation of spring coefficients with Poisson ratio is
here neglected without loss of accuracy for
the range of aspect ratio of interest, namely
the spring functions described above are
evaluated independent of the value of Poisson
ratio.

0.15

kx
D
= 1.828
E
B
k bx
D
= 0.669 + 0.129
EB
B
k
D
= 1.106 + 0.227
2
EB
B

0.15

kx
D
1.82
E
B -0.1495

y = 1.8283x
22
0.9483
RR ==0.9483

(35)
(36)
(37)

Calibration of dynamic springs

The fitted expressions and standard


deviation of the results is shown in Figures
11a-c for the lateral distributed, base concen-

For the calibration of the dynamic spring functions, the flexibility approach described above was also implemented here according to

379

lues of normalized frequency, and successively decreases for 0>1. Finally, for 0>2, it
attains values approximately zero, indicating
that no energy is being radiated towards the
far-field as a result of the rocking mechanism
along the shaft of the caisson. The rotational
attenuation coefficient is also found to increase in proportion to the aspect ratio (D/B), a
behavior that may be approximated by the
following expression:

which unit force and moment sinusoidal functions were applied at the top of the foundation to evaluate the dynamic response. Successively, the amplitude and phase difference
of displacements and rotations was measured, and the response was expressed by
means of complex functions that were next
used to estimate the complex springs.
Referring to equation (17), according to
which the dynamic springs are defined
as K * = K stat k '(ao ) + iaoC(ao ) , the variation of the
stiffness coefficient k(a0) and normalized
damping parameter C(a0) as a function of the
dimensionless frequency a0 is shown in Figures 12 a-f. The numerically derived variation
is shown to fluctuate with frequency, a behavior most probably attributed to local resonances within the truncated numerical domain
due to the finite dimensions of the model, as
well as potential scattered energy from the
absorbing far-field boundaries.
Nonetheless, for the range of aspect ratios
(D/B) investigated here, the variation of spring
coefficients with dimensionless frequency was
approximated by the following expressions,
which based on successive comparison with
numerical analyses were shown to yield satisfactory results in approximating the response
of caisson foundations subjected to dynamic
loading:
'
k x' = 1 0.1ao , k bx
= 1 , k' = 1 0.225ao
Cx 1.85ao
=
E 1.85

ao < 1
ao > 1

Cbx 0.6ao
=
EB 0.36

0.21( DB ) ao

C
= 0.21( DB ) (2 ao )
EB 2
0

ao < 1
1 < ao < 2
ao > 2

The validity of the approximate expressions


derived above is evaluated in the ensuing by
comparison of the analytically predicted response computed by means of the fitted expressions to the numerically predicted response
of the foundation-soil system.
KINEMATIC SSI INTERACTION

The inability of a stiff foundation to comply to


the deformation field imposed by the soil in
the free-field subjected to seismic incident waves and the resulting motion incompatibility is
translated into forces and moments being
applied on the foundation. In turn, the rigidly
responding foundation causes filtering of the
far-field motion, a phenomenon that materializes for wavelengths comparable or shorter
than the dimensions of the foundation. This
effect referred to as kinematic interaction and
schematically depicted in Figure 13, may be
expressed in terms of the following transfer
functions that illustrate the foundation
response resulting from the imposed free field
motion (i.e. the soil column deformation in absence of the foundation):

ao < 0.6
ao > 0.6

It should be also noted here that unlike the


variation of attenuation coefficients for the
translational distributed and base springs, the
corresponding values of the distributed rotational springs C are shown to yield negative
values (Figure 12f), an effect that while appearing similar to energy supply into the system,
it indicates that the waves produced by the
side shear resistance are out of phase with
those produced by other resisting mechanisms. As a result, the wavefield produced by
the lateral distributed shear resistance (i.e. the
mechanism corresponding to the lateral rotational springs) destructively interferes with the
wavefield produced by the translational
mechanisms, thus obstructing the energy radiation away from the system.
The distributed rotational attenuation
coefficient C increases with frequency for va-

Hu (ao ) =
H (ao ) =

ut
uff

B
uff

(38)
(39)

In order to evaluate the analytical


expressions for the transfer functions
describing the foundation response subjected
to seismic incident waves, consider a sinusoidal vertically propagating anti -plane shear
wave (SH) incident from the underlying
halfspace. The solution of the far-field displacement field in this case is given by:

380

1.5

1.5

k x

1.5

k xb

1.0

1.0

1.0

0.5

0.5

0.5

0
2.0

1.0

Cx E

0.4

0.5

-1.0

0.2
0

-0.5

0.6

1.0

EB
Cxb

0.8

1.5

-1.5

C E B 2
0

Fig. 12 Variation of numerically estimated values and fitted functions of dynamic stiffness and radiation
damping coefficients in Equation (17) as a function of the dimensionless frequency a0

uff ( z ) = uff cos(2

) = uff cos(ao

z
)
B

(40)

ut*

where z is the depth from the surface, is the


wavelength of the propagating sinusoid, and
a0 is the corresponding dimensionless frequency.

(( k

*
x

D2
*
*
D =
m 2 D + kbx
+ * k x* m 2
kbx
2

B
D
D
*
....uff* k x*
sin(ao ) + kbx
cos(ao )
B
B
ao

(41)

Similarly, requiring moment equilibrium at the


top of the foundation results in the following
expression:
3
*
2 D
2
+ ...
k x m

*
2 D
*
*
3
k bx D +
u k x m
2

*
2
*
*
... + kbx D + k D + kb
*
t


Jc

B 2

D
D
D
k x* 1 cos(ao ) ao sin(ao ) ...
B
B
B
ao

D
D
*
*
*
uff ... k bx D cos(ao ) k 1 cos(ao ) ...
(42)
B
B

... k * ao sin(a D )

b
o

B
B


mu
c

Equations (41) and (42) may be successively


expressed in a matrix form as follows:

Fig. 13 Schematic illustration of kinematic soilstructure interaction due to motion


incompatibility between the far-field and
the rigid foundation

*
K xx

EB
K xr*
2
EB

For the configuration shown in Figure 13,


equilibrium requirement of forces in the
horizontal direction results in the following expression:

Veff*
K xr* *

t
EB 2 = EB 2
B
*

K rr* * Meff
3
3
EB
EB

(43)

where the effective force vector corresponds


to the following quantities:

381

response, the model does predict the resonant and destructive interference frequencies
(i.e. the maxima and minima) of the corresponding transfer function within acceptable
range of accuracy. It should be also noted at
this point that the values predicted by 3-spring
Winkler model are for the most part conservative, namely the predictions represent less
pronounced reduction in the translational
motion and higher induced rocking motions.
Finally, the analytical transfer functions obtained by means of the model described above
as well as the results of 3D finite element
simulations are shown to be in the bounded
range predicted by means of the shallow
foundation theory (Elsabee and Morray, 1977)
and the pile kinematic response as formulated
by Gazetas et al (1993); this result implies
that the proposed model is indeed an
improvement to the existing models employed
for the simulation of rigid deep foundations
subjected to seismic motion, and by means of
a procedure that substantially reduces the
computational effort, the model can be used
to approximately predict the foundation
response subjected to transient seismic
excitation given the aspect ratio of the
foundation D/B.

*
Veff*
uff* k x* 1
D kbx
D
sin(
)
cos(ao )
=
a
+
(44)

o
2
B E ao
B EB
B
EB

D
)
2 1 cos(ao
*
B
kx 1

E a

o

D
D

a
sin(ao )
o

B
B

*
*
k bx D
D
Meff
uff*

...
cos(
)
a

=
o
(45)
EB B
B
B
EB 3

k*
D
1 cos(ao )
...
2
B
EB

kb*
D
ao sin(ao )
...

B
EB 3

By definition of the transfer functions given by


equations (38) and (39), the following
expression is derived for the estimation of the
foundation response upon the incidence of a
sinusoidal wave resulting in free-field response of amplitude uff:
ut*

Hu (ao ) uff
H (a ) = *
o B
uff

K*
xx
EB
= K*
xr
EB 2

K xr*

EB 2
K rr*

EB 3

Veff*

*
EBuff
M*
2eff *
EB uff

(46)

Three dimensional finite element numerical simulations were here conducted to evaluate the target transfer functions Hu and H
for the case of vertically propagating SH waves. For the numerical representation of the
problem, the far-field motion was applied in by
means of effective forcing functions in the
interior of the truncated computational domain, while the potentially scattered wavefield
was verified to be almost entirely absorbed by
the sponge boundaries. A comparison with
the values derived by means of the analytical
expression using both fitted and numerically
derived (unfitted) spring constants is shown in
Figures 14a and 14b for D/B =4.
Results obtained in this section illustrate
that the analytical solution for the kinematic
response of rigid intermediate embedded
foundations obtained by means of the 3spring model, allows the kinematic interaction
effects of pier foundation elements to be captured within an acceptable degree of accuracy, considering the substantial reduction in
computational time. In particular, while the 3D
finite element simulations do not capture
quantitatively the amplitude of the rocking

1.6
Flexible pile SSI
Embedded foundations
Spring interpolated values
Fitted spring functions
3D Simulations

Hu

1.2
0.8
0.4
0

a0
0

0.75

H
0.50
0.25

a0
0

Fig. 14 Displacement and rotational transfer function for kinematic SSI of a weightless
foundation with aspect ration (D/B=4)
Transient response of caisson

The analytical expressions developed above


for the displacement and rotation transfer
functions from the far-field deformation cau-

382

numerical modeling aspects of the problem


that may strongly affect the computed
response such as the selection of proper
integration scheme and time-step, finite
element model discretization and effectively
absorbing boundary conditions. Nonetheless,
while the dominant frequency in both translational and rocking motion response evaluated by means of the analytical model is
found to be in excellent agreement with the
numerical results, the maximum translation
predicted by the analytical model is 0.37m
(which is higher compared to the numericallyobtained value umax = 0.23m), and the pier
rotation is predicted to be 0.34rad by means
of the analytical model (higher than the corresponding value of 0.22rad obtained by means
of the numerical simulations). It should be
noted though that the results of the example
shown here are stemming from the overprediction of the system stiffness by means of
the four mechanisms of resistance that do not
capture the complete physical problem.
Nonetheless, the approximate model is expected to yield conservative results for the predicted translations and rotations compared to
3D finite element simulations and does provide a better approximation to the existing
simplified models available for the evaluation
of kinematic response of intermediate rigid
embedded foundations.

sed by vertically propagating SH seismic


waves to the rigid pier translation and rotation
may be implemented at minimal computational cost in a computer script and used to
estimate the motions corresponding to any
transient loading by means of Fourier reconstruction of the response signal, provided that
the medium of propagation is linear elastic or
moderately nonlinear. Figures 15a and 15b
show an example of seismic displacement
time history prescribed at the base of the numerical model and its Fourier transform. For
the numerical simulations, the far-field (1D)
response is initially computed and imposed at
the corresponding location of the numerical
domain (i.e. at the soil-sponge boundary interface) in the form of effective forces, and
results are compared in the ensuing with the
corresponding ones obtained by means of
Fourier analysis using the transfer functions
described in equations (38) and (39) and the
dynamic fitted springs developed in the
foregoing sections.
0.50
0.25
0
-0.25
-0.50

10

0.3

0.4

0.2

0.3

3D Finite elements
3-spring model

0.2

0.2

-0.4

0.1

-0.2

0.1

0.4

10

0.2

-0.2

0.1
0

0.3

-0.4

0.4

3D Finite elements
3-spring model

0.2

Fig. 15 Displacement time history and corresponding Fourier spectrum prescribed at the
base of the numerical domain for the illustration of the model performance in transient excitation

10

Fig. 16 Comparison of translation and rotation of


caisson foundation subjected to transient
motion between the analytically predicted
and numerically obtained results.

The comparison between the analyticallypredicted and numerically-evaluated response


is shown in Figures 16a and 16b. Note that
the numerically-obtained time histories have
been shifted in time to account for the wave
propagation duration of the excitation traveling
from the far-field to the pier. Results presented above readily illustrate that that the model
is able to capture the response of pier within
an acceptable degree of accuracy while substantially reducing the computational time and

CONCLUSIONS

We have described the development of an


analytical model for the prediction of the response of rigid cylindrical caisson foundations
characterized by aspect ratios D/B = 2-6 and
embedded in linear elastic soil media, using a

383

may be even further extended to capture the


nonlinear soil behavior by implementing
nonlinear spring elements that represent the
macroscopic response of the material as a
function of depth. It should be noted also that
in particular for the representation of
kinematic interaction, a first approximation to
the nonlinear response of the soil-foundation
system (provided that the foundation material
is responding within the linear elastic range)
could be obtained by means of equivalent
linear analyses in the far-field, successively to
be used as effective forcing far-field response
function at the base and soil-foundation
interface for the reduced stiffness and
material damping evaluated at convergence of
the algorithm. Based on the developed
computational platform, analytical formulations may finally be developed for the motionresponse transfer functions of the pier subjected to horizontally propagating coherent SH
waves or for the response of intermediate
rigid foundations of various cross-sectional
geometries.

simple Winkler spring model with four springs.


Results illustrated in this paper show that for
the range of aspect ratio of interest (D/B = 26), the effect of base rotational spring is
indeed negligible and therefore, a simplified
three spring model may be used instead to
capture the pier response. Based on this approximation, expressions were developed for
the three springs as a function of both the
aspect ratio (D/B) and the dimensionless loading frequency since sensitivity analyses
showed that the effects of Poisson ratio may
be neglected in the evaluation of the
aforementioned functions.
Analytical expressions were developed for
the global foundation stiffness matrix and for
the transfer functions of kinematic interaction
effects, for the prediction of forced vibration
and seismic transient response correspondingly. The theoretical values of free-field/pier
response transfer functions for translational
and rocking motion resulting from the seismic
excitation were compared to the values
obtained by means of 3D finite element
simulations, and despite the fact that the
proposed formulation does not simulate the
pier response exactly -attributed to the
complex load transfer mechanisms applied at
the soil-foundation interface that cannot be
captured by the simplified 3-spring proposed
model- it may be applied to capture the
important response parameters, namely the
frequency content and evolution of timehistory variation.
It should be noted here that while the
Winkler springs developed in this study are
applicable for linear elastic medium with no
material damping, the effect of material damping can be easily accounted for by means of
the elasticity-visco-elasticity correspondence
principle, namely the use of a complex modulus E*=E(1+2i), where is the material damping ratio.
Furthermore, while the formulation presented in this paper is not capable of capturing
the separation between the soil and pier interface, modified Winkler springs that include a
stiffness element, a damper and a Coulomb
friction element with low tension resistance
may be implemented to account for the separation at the soil-foundation interface, and
the variability in soil strength subjected to
compression versus extension. The approach

REFERENCES
Assimaki D., Chatzigiannelis I., Gerolumos N.
and Gazetas G. (2001).Lateral response
of caisson foundations, Proc. 4th National
Conference
on
Geotechnical
and
Geoenvironmental Engineering, Athens,
Greece, June (in Greek)
Dominguez, J. (1978) Dynamic stiffness of
rectangular foundations, Research Rep.
R78-20, MIT
Davidson, H.L. (1982) Laterally Loaded Drilled
Pier Research, Research Rep. EL-2197,
EPRI
Elsabee, F and Morray, J.P. (1977) Dynamic
behavior of embedded foundations,
Research Rep. R77-33, MIT
Gazetas, G. (1983) Analysis of Machine
Foundation Vibrations: state of the art, J.
Soil Dynamics and Earthquake Engng., 2,
2-42.
Gazetas, G., Fan, K. and Kynia, A. (1993)
Dynamic response of pile groups with
different configurations, J. Soil Dynamics
and Earthquake Engng., 12, pp. 239-257

384

Gerolymos N. and Gazetas G. (2006) Winkler


Model for Lateral Response of Rigid
Caisson Foundations in Linear Soil, Soil
Dynamics & Earthquake Engineering, Vol.
26, No. 5, pp 347-361 2006.

case. J. Eng Mech Div, ASCE, 104(4):


953-9
ONeil, M.W. and Reese L.C. (1999) Drilled
shaft: Construction procedures and design
methods, Research Rep. FHWA-IF-99-025

Gerolymos N. and Gazetas G. (2006)


Development of Winkler Model for Static
and Dynamic Response of Caisson
Foundations with Soil and Interface
Nonlinearities,
Soil
Dynamics
&
Earthquake Engineering, Vol. 26, No. 5,
pp 363-376, 2006.

Prevost, J.H. (1983) Dynaflow. Princeton


University, Princeton, NJ 08544
Tajirian, F.F. and Tabatabaie, M. (1985)
Vibration analysis of foundations in
layered media. In: Gazetas, G. and Selig,
E.T., editors. Vibration problems in
geotechnical engineering, ASCE, NY

Gerolymos N. and Gazetas G. (2006) Static


and Dynamic Response of Massive
Caisson Foundations with Soil and
Interface Nonlinearities-Validation and
Results, Soil Dynamics & Earthquake
Engineering, Vol. 26, No. 5, pp, 377-394,
2006 .

Tassoulas, J.L. (1981) Elements for the


numerical analysis of wave motion in
layered media, Research Rep. R81-2, MIT
Veletsos, A.S. and Verbic, B. (1974) Basic
response functions for elastic foundations,
J.Engrg. Mech. Div, ASCE, 100, EM2, 189

Hibbit, Karlssson & Sorenson (2006)


ABAQUS users manual, Pawtucket, R.I.

Veletsos, A.S. and Wei, Y.T. (1971) Lateral


and rocking vibrations of footings, J. Soil
Mech. Found. Div., ASCE, 97, SM9, 1227.

Kausel, E. (1974) Forced vibrations of circular


foundations on layered media, Research
Rep. R74-11, MIT

Wolf, J.P. (1997) Spring-Dashpot-Mass


models
for
foundation
Vibrations,
Earthquake
Engg.
and
Structural
Dynamics, Vol. 26, pp. 931-949.

Kausel E and Rosset JM (1975) Dynamic


stiffness of circular foundations. J. Eng
Mech Div, ASCE, 101(6): 770-85
Lam, I.P. and Chaudhury, D. (1997) Modeling
of drilled shafts for seismic design
NCEER reprt for task no II-2D-2.5,
NCEER, SUNY Buffalo.
Mayne, P.W., Kulhawy, F.H., and Trautmann,
C.H. (1992) Experimental study of
undrained lateral and moment behavior of
drilled shafts during static and cyclic
loading, Report TR-100331, Electric
Power Research Institute, Palo Alto, 350 p.
Mita, A. and Luco, J. E. (1989) Dynamic
response of a square foundation
embedded in an elastic halfspace, Soil
Dynamics and earthquake engineering,
8(2), 54-67.
Novak, M. and Beredugo, Y.O. (1972),
"Vertical Vibration of Embedded Footings,"
J. Soil Mech. and Found. Div. ASCE, 98,
No. SM12, pp. 1291-1310
Novak M, Nogami T and Aboul-Ella F. (1978)
Dynamic soil reactions for plane strain

385

Seismic zonation, vulnerability assessment and loss scenarios in


Thessaloniki
K. Pitilakis1, A. Anastasiadis1, K. Kakderi1, S. Argyroudis1, M. Alexoudi1
1

Aristotle University of Thessaloniki, Greece

Abstract
The paper presents the application in Thessaloniki Greece of a general and modular
methodology for constructing earthquake-risk scenarios that takes into consideration the
inventory, typology and vulnerability characteristics of the elements at risk, as well as the
seismic hazard, geotechnical characterization and a detailed site response of the main soil
formations. Given the spatial distribution of the characteristics of earthquake motion, loss
scenarios for buildings, lifelines and transportation systems are produced using inventory
data and adequate fragility curves. Herein, examples of estimated earthquake losses for the
water and transportation (bridges) system in Thessaloniki as well as for the building stock in
the central part of the city are presented for the seismic scenario with mean return period
Tm=475 years. Furthermore, the estimation of casualties and direct economic losses is
performed using appropriate models taking into consideration the features of the local
construction practice and experience from past earthquakes. Finally, the effects of various
mitigation schemes that could be incorporated for the reduction of potential losses are
investigated for the current building stock. The previous applications reveal the importance
of site specific seismic response analysis (i.e. microzonation study) for the vulnerability
assessment and the definition of efficient mitigation strategies and policies for pre and post
earthquake actions, in order to reduce the expected consequences from different
earthquake events.

zonation studies in Thessaloniki for a specific


seismic scenario and we highlight their
importance for the assessment of adequate
loss scenarios for various elements at risk, the
selection of effective retrofitting schemes of
existing structures and finally for the efficient
risk management.

INTRODUCTION

High seismicity in Greece in combination


with an extensive urbanisation and a relatively
fast increase of value in real estate and
infrastructure, have caused earthquake risk to
increase in recent years, despite progress in
construction and earthquake engineering
standards. Thus it is imperative to define and
implement efficient mitigation strategies and
policies for pre and post earthquake actions in
order to reduce the expected consequences
from different earthquake events. The latter
could only be based on adequate loss
scenarios which take into consideration the
inventory,
typology
and
vulnerability
characteristics of the elements at risk, as well
as
the
seismic
hazard,
geotechnical
characterization and site response of the main
soil formations for different seismic scenarios.
In the present study we present the seismic
hazard, site characterization and seismic

METHODOLOGY

The discrete steps of the loss modelling


procedure that was followed in the case of
Thessaloniki are illustrated in Fig 1. Loss
estimates including physical damage, casualties
and direct economic losses depend on the
existing inventory and typology classification of
the elements at risk, the vulnerability and
casualty models to be used and the alternative
mitigation actions to be implemented. The level
of seismic input motion is defined on the basis
of a Microzonation study in order to take into
account the effects of the seismic response of

386

the soil formations. Finally, the related


uncertainties due to inventory data, seismic
hazard analysis, vulnerability and casualty
models should be taken into consideration.

(Papazachos et al., 1979). It is a Quaternary


formation, tectonically re-activated during the
Neocene-Quaternary, in contact with the
Jurassic subduction zone of Vardaris ocean
under the continental Servomakedonian zone,
with nearly vertical normal faults (Mercier,
1968). It is classified as a tectonically active
region, with the presence of several active
faults. During the Neotectonic period (Lower
Neogene and mostly Quaternary period)
extensive tectonic submersion zones and sinks
have been happened creating the present
active normal fault system with an E-W
direction.
The city of Thessaloniki has experienced
several destructive earthquakes since its
foundation. The available historical and
instrumental data indicate three discrete
periods of high seismic activity near the city (7th
AD, 15th -18th, and 20th centuries). Despite the
relative accuracy of the historic earthquakes,
referring mostly to the oldest ones and
especially to their epicenter geographic
position, the earthquakes practically affecting
the city occur from Axios, Mygdonias and
Anthemountas graben zones.
The latest major earthquake registering a
magnitude M of 6.5, struck the city in 1978. Its
epicentre was located about 30 km east of the
city. It resulted in 50 deaths and severely
damaged a significant number of buildings. The
public repair cost of the particular residences
was of the order of 250 million US dollars
(1978). The total economic damage was
certainly very high.

SEISMIC SCENARIOS MICROZONATION


STUDY

In Thessaloniki a detailed Microzonation


study has been conducted for three different
mean return periods Tm=100, 500 and 1000
years. The study is based on the results of a
probabilistic seismic hazard analysis using
recently obtained data regarding the seismicity,
the corresponding seismic zones and the
seismic faults in the greater area (RISKUE,
2001-2004, SRMLIFE, 2003-2006). Site effects
are calculated performing a great number of 1D
linear equivalent response analyses in order to
take into account the influence of geotechnical
characteristics and dynamic properties of the
main soil formations on expected seismic
ground motion. The analysis is performed with
five different scaled real accelerograms
(representing bed rock conditions) selected
according to seismic hazard study.
Geological and Morphological Setting

Thessaloniki and its suburbs are located


along Thermaikos Gulf and have a rather
smooth topographical relief. The geology of the
broader area of Thessaloniki is shown in Fig 2.
The metropolitan area of Thessaloniki is
situated on three (3) main large-scale geology
structures, oriented in NW-SE direction. The
first formation includes the metamorphic
substratum consisting of gneiss, epigneiss, and
green schists, which are surficial near the city at
the N-NE border of the urban area. These
crystalline rocks constitute the bedrock
basement beneath the city reaching a depth of
150-300m near the coastline in W-WS direction.
The second formation is composed by alluvial
deposits mainly of the Neogene period. In this
geological structure the red silty clay, series are
dominant, covering the bedrock basement
beneath the city. Finally, recent deposits
consisting of Holocene clay-sand-debris
geological materials compose the third surficial
unit.

Geotechnical Zonation
characteristics

and

Dynamic

soil

Reliable modeling and evaluation of site


effects is the key problem in seismic
microzonation especially in high seismicity
areas, with complex geology and highly
heterogeneous soils. Additionally, risk reduction
studies require the detailed, spatial distribution
of specific parameters (e.g. PGA, PGV, PGD) in
a broader area, for different seismic scenarios.
Hence, a detailed model of the surface
geology and geotechnical characteristics,
properly oriented for site effect studies, was
generated for the city of Thessaloniki. The
resulted geotechnical map (Anastasiadis et al.,
2001) was based on numerous data provided
by geotechnical investigations (boreholes,
CPTs, water wells), geophysical surveys (cross

Seismicity and tectonics

Thessaloniki is located close to one of the


most active seismotectonic zones in Europe

387

PGA(gs) at outcropping bedrock conditions for


the area in interest, stemming from seismic
hazard analysis for 475 years return period
(SRM-LIFE Report , Papaioannou 2004).
On the basis of the earthquake scenario
with 10% probability of exceedance in 50 years
(mean return period of 475 years), according to
the results of seismic hazard study, the
characteristics of the calculated seismic ground
motions at the free surface, in terms of peak
acceleration (PGA) and velocity (PGV) are
presented in Figs 7 and 8.

holes, down holes, surface seismics),


microtremors
measurements,
classical
geotechnical and special soil dynamic tests
(resonant column, cyclic triaxial) (Pitilakis et al.,
1992, Pitilakis and Anastasiadis, 1998,
Raptakis et al., 1994a, Raptakis et al., 1994b,
Raptakis 1995). The thematic GIS maps in Fig
3 illustrate the spatial distribution and the
thickness of the main soil formations. In total
nine (9) different soil formations are needed to
fully describe the subsoil conditions in the city
(Table 1). These maps resulted from the
synthesis of all available data and depict the
thickness of the A and C soil formations, the
depth of the stiff clayey formations (E and F)
and the depth of the rock basement (formation
G) with respect to the regions where they are
predominant. The variation of the thickness and
the depth shows a certain irregularity of the
subsoil structure beneath the urban area of the
city.
Degradation curves of shear modulus G/Go and damping ratio Ds-, determined for all
nine-soil formations from an extended
laboratory including resonant column and cyclic
triaxial tests (Pitilakis et al., 1992, Pitilakis and
Anastasiadis, 1998, Anastasiadis, 1994) are
depicted in Fig 4.

VULNERABILITY ASSESSMENT AND LOSS


SCENARIOS

Seismic hazard for the vulnerability analysis


and risk assessment of buildings and lifelines
should be specified according to the precise
needs for the particular networks as well as the
most adequate models used to describe
vulnerability and fragility relationships. Moreover
due to the spatial extend of lifeline systems,
spatial variability of ground motion considering
the local soil conditions is of great importance
(Pitilakis et al., 2005). Thus, loss scenarios for
buildings, lifelines and transportation systems
could be constructed on the basis of site
specific seismic hazard analysis using available
inventory data and adequate fragility curves.
Herein, the assessment of potential earthquake
losses is performed for the water system in
Thessaloniki, the bridges and the buildings in
the central part of the city based on the results
from the Microzonation Study for the basic
scenario with mean return period Tm=475
years.

Site Response Analysis

The geotechnical map together with


thematic GIS maps describing the spatial
distribution and the thickness of each soil
formation combined with detailed 2D design
cross sections, were used to develop adequate
1D soil profiles at 520 sites in a quadratic grid
ranging from 1000m x 1000m to 250m x 250m
(Fig 5).
1D-EQL soil response analyses (Schnabel
et al., 1972) are conducted using as input
motion recorded acceleration time histories of
five earthquakes, which were found to have
characteristics that cover satisfactorily the
intensity and frequency content of the
earthquake scenario bedrock motion in the area
of Thessaloniki. The recorded input motions
were scaled properly according to the seismic
hazard results referring at three scenarios
(recurrence period of 100, 475 and 950 years)
of the urban area and the similarity of the
average acceleration response spectrum values
with the scenario spectrum on hard soil or
outcropping bedrock conditions. Fig 6 depicts
representative results in terms of mean

Vulnerability Assessment
Water System

of

Thessalonikis

Thessalonikis potable water system is


comprised of about 1351 km of pipes. The
current inventory database includes several
attributes such as location, diameter, material,
age, operating area, supplied tank, type, depth,
length, joint type and history of failures. It is
worth mentioning that important information
such as material, diameters and ages of water
pipes is missing in many cases; for instance
about 42.5% of pipes have unknown diameter
and large areas have pipes of unknown
construction
date
and
materials.
The
vulnerability assessment of potable water pipes
is based on the estimation of the expected

388

conditions and the proximity of the seismic


source (ex. southeast part). For instance, in the
northwest part of the city the soil is described
as deep soft alluvium deposits, sandy-silty clays
to clayey sands-silts, with low strength and high
compressibility, (category C and D in EC8); thus
the seismic motion presents a stronger
amplification at longer periods.

Repair Rate per pipe km (RR/km). Expected


damages (leaks and/or breaks) caused by wave
propagation are estimated using O Rourke &
Ayala (1993) fragility relation proposed by
HAZUS (NIBS, 2004). The seismic loading is
described in terms of peak ground velocity
(PGV), because it is directly proportional to the
ground strains. Fig 9 presents the spatial
distribution of Thessalonikis water system
damages for the 475 years scenario.
A total number of 44 repairs are anticipated.
Assuming that in areas where ground behavior
is dominating by wave propagation, 80% of
pipeline damages should be considered as
leaks and the other 20% as breaks, a number
of 9 breaks and 35 leaks are estimated for
Thessalonikis potable waterpipe system. The
number and location of the damages is related
to the spatial distribution of seismic motion of
the specific scenario as well as the individual
characteristics of the examined elements. The
direct economical cost that is related to the
damaged water pipes replacements according
to Greek experience ranges between 66.000118.800 Euro.

Vulnerability Assessment and Loss Estimates


for Buildings

Loss estimates are performed for the central


part of the city which covers about 40% of the
municipality of Thessaloniki. A detailed building
inventory based on the results of the RISK-UE
project (RISK-UE, 2001-2004) has been used
[5,032 reinforced concrete (R/C) and masonry
(URM) buildings, approximately 100,000
inhabitants]. Population data and land uses
from the 2001 Census have also been
incorporated (SRMLIFE, 2003-2006). The
reference unit of the inventory is the building
block; the same has been used for the
vulnerability analysis and loss estimations.
The classification of the buildings is based
on the Building Typology Matrix of the RISKUE
project (Kappos et al., 2004), according to the
material, structure type, height and code level.
Appropriate vulnerability classes can be
specified based on the discrete features of each
building. The prevailing typology in the area is
R/C buildings designed to old codes.
The vulnerability analysis of R/C buildings is
performed based on fragility curves (in terms of
PGA) that have been developed using a hybrid
technique combining analytical results and
statistical data (Kappos et al 2006). Six damage
states (DS) are defined: no damage (DS0),
slight (DS1), moderate (DS2), substantial to
heavy (DS3), very heavy (DS4) and destruction
(DS5). According to empirical estimations for
Greece it seems that only 10% of the totally
damaged buildings are expected to fully
collapse. Moreover, fragility curves (in terms of
both PGA and Sd) for masonry structures that
were developed for all typologies common in
Greece were used in the present application
(Penelis et al. 2002).
Fig 11 illustrates the resulted damage
distribution in terms of the predicted threecolour building tagging (OASP, 1999). Relative
results are also given in Table 2. The
correspondence between tag colors and
damage states was assumed as follows:

Vulnerability Assessment of Thessalonikis


Transportation System (Bridges)

The current inventory for the roadway


network in the Metropolitan area of Thessaloniki
includes about 80 bridges. The classification of
bridges is based on the number of spans
(single or multiple), the seismic code level (low
or upgraded), the column bent type (single or
multiple) and the span continuity (continuous or
simple support). The vulnerability analysis is
performed based on the fragility curves that are
provided by HAZUS (NIBS, 2004), while the
input earthquake hazard scenario was obtained
from the Microzonation study and is referred to
the mean spectral acceleration at T=1.0sec. Fig
10 presents the estimated worst probable
damage state (i.e. exceeding probability >50%)
for each bridge.
The application of the fragility model shows
that the majority of bridges will respond
satisfactory, but there are still few bridges,
which are expected to sustain serious damage
for the specific seismic hazard scenario. This is
due to the higher vulnerability of these bridges
(single column, simple support bridges and
inadequate seismic design) and the higher
values of the expected surface acceleration.
The latter is attributed to the local soil

389

Green: DS0 & DS1 (light mostly non-structural


damage).
Yellow: DS2 & DS3 (repairable structural or
non-structural damage).
Red: DS4 & DS5 (severe structural damage
that would mostly be demolished).
It is seen that the most heavily damaged
typologies are low-code R/C buildings with
infilled frame systems. The latter is anticipated,
whereas the particularly high vulnerability of
low-rise old R/C buildings maybe attributed to
their location with respect to the specific
considered scenario and their own (absolute)
vulnerability.
Human casualties are estimated using the
Coburn and Spence, 2002 model. The different
M parameters (M1-M5) introduced in the
relationship are estimated per building block. A
factor of 0.01 is applied to the total number of
destructed buildings in order to define the
number of total collapses (i.e. collapses that
potentially involve casualties). This value is
proved to be reasonable when examining the
consequences of the 1999 Athens earthquake
(Pomonis, 2002), and it is in agreement with
HAZUS (NIBS, 2004) loss estimation approach.
The application of the model is performed
assuming two different times of the earthquake
event, namely at 12:00h and 24:00h. Finally,
direct economic losses are estimated using the
respective loss indices of the fragility models
used (Pitilakis et al, 2004), while an average
replacement cost of 700 /m2 is assumed.
Referring to the specific urban area
(100,000 inhabitants), a rate of 5 and 6 (12:00
and 24:00 hours scenario respectively) dead
people are estimated for the scenario of 475
years mean return period. It seems that the
critical parameter for human casualties
estimates is the total number of buildings that
fully or partially collapse entrapping people.
Finally, estimated total costs of repair are
approximately equal to 382 million euros. The
distribution of cost is, of course, consistent with
(and conditional on) the distribution of the
degree of damage in terms of the total floor
area being in each damage state per building
block.

uncertainties inherent in the seismic hazard and


the systems response, are systematically
incorporated in any decision of intervention
schemes to be applied for the reduction of
earthquake risk (Park, 2004). Several factors
like structural performance and/or direct
structural cost, life losses and secondary
economic losses, should be taken into
consideration.
Three mitigation options are identified for
the R/C buildings in this study for Thessaloniki.
The basic assumption that a retrofitted building
corresponds, in analytical terms, to another
upgraded building type from the existing
categories is made. Thus, already available
fragility curves (of an upper class) are used to
perform the vulnerability assessment of the
retrofitted buildings. The aforementioned
mitigation options are outlined in Table 3.
Expected building damages are provided in
Table 2 prior and after retrofitting. There is a
considerable reduction of collapsed buildings
(almost 40%). Estimated human casualties are
reduced to 3/5, 1/2 and 3/4 dead people
(12:00/24:00 hours scenario) for mitigation
options 1, 2 and 3 respectively.
Associated costs for different mitigation
options include the cost of structural
intervention as well as a minimum percentage
of intervention to existing structural and/or
architectural elements of the retrofitted
buildings (Stylianidis, 2006). They are provided
as a percentage of the construction cost for
new buildings. In the framework of the present
study appropriate costs have been estimated
according to each mitigation option and the
height class of the buildings as a percentage of
the structural system construction cost. Total
repair costs for mitigation options 1, 2 and 3 are
estimated to be approximately equal to 325,
174 and 314 million euros respectively (15, 54
and 18% reduction of the initial prior retrofitting
total repair cost).
Comparative study of the three different
mitigation options examined, shows that the
first and third intervention schemes (upgrading
of all low-code, R/C frame buildings
corresponding to 5% of the building stock)
result in similar effects concerning the potential
physical damage to buildings, repair costs and
estimated human casualties, while referring to
the same number of retrofitted buildings. More
specifically, the third mitigation option results in

Retrofit of Existing Structures

Seismic rehabilitation of structures is


certainly the best way to reduce seismic losses.
Multiple conflicting criteria and various

390

the 475 years scenario, examples of the


assessment of potential earthquake losses are
presented for the water system, the bridges and
the buildings in the central part of the city.
On the basis of a decision analysis for
identification of the optimal target reliability level
for rehabilitation of a building structure against
seismic hazard as well as the best solution
among
several
alternative
rehabilitation
schemes, different mitigation actions could be
identified and implemented. In the case of
Thessaloniki, the effects of three mitigation
schemes that could be proposed for the
reduction of potential losses are investigated.
However the final decision is conditional to the
available budget and the desirable level of the
intervention and of course the anticipated level
of physical damage for the specific seismic
scenario.
Finally based also on the previous
applications, the importance of site specific
seismic response analysis (Microzonation
study), is revealed for the vulnerability
assessment and the definition of efficient
mitigation strategies and policies for pre and
post earthquake actions in order to reduce the
expected
consequences
from
different
earthquake events.

slightly better effects but the associated cost is


higher. On the other hand, the latter seems to
have a more realistic mitigation goal
compared to the first one. The second
retrofitting alternative [upgrading of all lowcode R/C buildings (both frames and dual
systems) corresponding to 74% of the building
stock] gives higher percentage of loss
reduction; thus a quite large and rather
prohibitive number of buildings are involved.
Conclusively, the first and third intervention
schemes could be considered as the most
appropriate for the area of the city examined
(central part of the city) where limitations of the
number of retrofitted buildings exist. It is pointed
out that the final decision on mitigation action(s)
and level of intervention should be based on
estimated results in relation to the reduction of
potential losses, direct and indirect costs
estimates for different mitigation actions and
expected economic losses. However the final
decision is conditional to the available budget
and the desirable level of the intervention.
Finally, it is pointed out that a rigorous and
effective risk management plan should be
based on the comparative consideration of
estimated losses through the assessment of the
relative effects of different mitigation actions.
CONCLUSIONS

REFERENCES

In the present study the application of a


general and modular methodology for
constructing earthquake-risk scenarios is
presented for the city of Thessaloniki in Greece.
Seismic risk scenarios take into consideration
the inventory, typology and vulnerability
characteristics of different elements at risk, as
well as the seismic hazard, geotechnical
characterization and site response of the main
soil formations for different seismic scenarios.
Thus, vulnerability and loss estimates for
buildings, lifelines and transportation systems
as well as the assessment of casualties and
direct economic losses are evaluated on the
basis of site specific seismic hazard analysis
using available inventory data and adequate
fragility curves.
Herein, a very short presentation of the site
characterization and seismic zonation for the
city of Thessaloniki is presented. A detailed
Microzonation study has been conducted for
three different mean return periods Tm=100,
500 and 1000 years. Based on the results for

Anastasiadis, A., (1994), Contribution to the


determination of the dynamic properties of
natural Greek soils, Ph.D. Thesis (in Greek),
Dep. of Civil Engineering, Aristotle University of
Thessaloniki.
Anastasiadis, A., Raptakis, D., Pitilakis K., (2001.),
Thessalonikis
Detailed
Microzoning:
Subsurface Structure as basis for Site Response
Analysis, Pure and Applied Geophysics
PAGEOPH, Vol. 158, No. 12, pp. 2597-2633.
Coburn, A. and Spence, R. (2002) Earthquake
Protection (2nd edition), J. Whiley and Sons Ltd.,
pp. 420.
IGME, (1978), Geological Map of Thessaloniki,
Institute of Geology and Mineral Exploration,
(scale - 1:5000).
Kappos, A., Panagopoulos, G., Panagiotopoulos,
Ch., and Penelis, Gr. (2006) A hybrid method for
the vulnerability assessment of R/C and URM
buildings, Bulletin of Earthquake Engineering,
Vol. 44, pp. 391-413.

391

Kappos, A.J., Panagiotopoulos, Ch., Panagopoulos


G., and Papadopoulos, El. (2004) WP04
Reinforced concrete buildings (Level I and II
analysis) RISK-UE Report, Thessaloniki.

Pitilakis, K., and Anastasiadis, A., (1998), Soil and


site characterization for seismic response
analysis, Proceeding of the XI ECEE, Paris 6-11
Sept. 1998, Inv.Lectures, pp.65-90.

Mercier, J.L., (1968), Etude geologique des zones


internes des Hellenides en Macedoine centrale.
Contribution a letude du metamorphisme et de
levolution magmatique des zones internes des
Hellenides , Ann.Geol. des Pays Hell., Vol. 20,
pp. 1-735.

Pitilakis, K., Kappos, A., Hatzigogos, Th.,


Anastasiadis, A., Anastasiadis, A., Alexoudi, M.,
Argyroudis, S., Penelis, G., Panagiotopoulos,
Ch.,
Panagopoulos,
G.,
Kakderi,
K.,
Papadopoulos, I., Dikas, N., (2004) Synthesis of
the application to Thessaloniki city, RISK-UE
Report, Thessaloniki.

National Institute of Building Sciences (NIBS) (2004)


Earthquake Loss Estimation Methodology
HAZUS 2004, Service Release 2: Technical
Manual, FEMA, Washington DC.

Pomonis, A. (2002) The Mount Parnitha (Athens)


Earthquake of September 7, 1999: A Disaster
Management Perspective, Natural Hazards, Vol.
27, pp. 171-199.

OASP (1999) Implementation Guidelines of the


Immediate
First
Level
Post-earthquake
Assessment of the Usability of Damaged
Buildings (in Greek), Booklet published by the
Earthquake
Protection
and
Planning
Organisation, Athens, Greece.

Raptakis, D., (1995), Contribution to the


determination of the geometry and the dynamic
characteristics of soil formations and their
seismic response, Ph.D. Thesis (in Greek), Dep.
of Civil Engineering, Aristotle University of
Thessaloniki.

O'Rourke, M., and Ayala, G., (1993), Pipeline


Damage due to Wave Propagation, Journal of
Geotechical Engineering, ASCE, Vol. 119, No.9.

Raptakis, D.G., Anastasiadis, A.J., Pitilakis, K.D. and


Lontzetidis, K.S., (1994a), Shear wave velocities
and damping of Greek natural soils, Proceeding
of the 10th ECEE Vienna, Austria, Vol.1, pp. 477482.

Papaioannou, Ch., (2004), Technical Report


Research Program SRM-LIFE, Seismic Hazard
Scenarios: Probabilistic Analysis of Seismic
Hazard, Ch. Papaioannou, Thessaloniki, 2004.

Raptakis, D.G., Karaolani, E., Pitilakis, K.,


Theodulidis, N., (1994b), Horizontal to vertical
spectral ratio and site effects: The case of a
downhole array in Thessaloniki (Greece),
Proceeding of XXIV General Assembly, ESC,
Athens, Vol.III, pp.1570-1578.

Papazachos, B.C., Moudrakis, D., Psilovikos, A. and


Leventakis, G., (1979) Surface fault traces and
fault plane solutions of the May-June 1978.
Major shocks in the Thessaloniki area, Greece,
Technophysics, Vol. 53, pp. 171-183.
Park, J. (2004) Development and application of
probabilistic decision support framework for
seismic rehabilitation of structural systems,
Thesis, Georgia Institute of Technology.

RISK-UE, (2001-2004) An Advanced Approach to


Earthquake Risk Scenarios with Applications to
Different European Towns Research Project,
European Commission, DG I2001-2004, CEC
Contract Number: EVK4-CT-2000-00014.

Penelis, Gr.G., Kappos, A.J., Stylianidis, K.C., and


Panagiotopoulos, C. (2002) 2nd level analysis
and vulnerability assessment of URM buildings,
International Conference Earthquake Loss
Estimation and Risk Reduction, Bucharest,
Romania.

Schnabel, P.B., Lysmer, J., and Seed, H.B., (1972),


SHAKE: A computer program for earthquake
response analysis of horizontally layered site,
Rep. No.EERC 72-12, Earthquake Engineering
Research Center.
SRMLIFE (2003-2006) Development of a global
methodology for the vulnerability assessment
and risk management of lifelines, infrastructures
and critical facilities. Application to the
metropolitan area of Thessaloniki Research
Project, General Secretariat for Research and
Technology, Greece.

Pitilakis, K., Alexoudi, A., Argyroudis, S., Monge, O.,


and Martin, C. (2005), "Chapter 9: Vulnerability
assessment of lifelines", C.S. Oliveira, A. Roca
and X. Goula ed. "Assessing and Managing
Earthquake Risk. Geo-Scientific and Engineering
Knowledge for Earthquake Risk mitigation:
Developments, Tools and Techniques", Springer
Publ.

Stylianidis, K. (2006), personal communication.

Pitilakis, K., Anastasiadis, A., Raptakis D., (1992),


Field and Laboratory Determination of Dynamic
Properties of Natural Soil Deposits, Proceeding
of the 10th WCEE, Madrid, Vol.5, pp.1275-1280.

392

Table 1. Dynamic properties of the main soil formations of Thessaloniki urban area. The values in brackets
specify the mean values of VS velocities and quality factors QS.
Formation

Description

(1)

(2)

B2
B3

Surficial

B1

C
D

Subbase

Artificial Fills, demolition


materials & debris parts
Very Stiff sandy-silty clays to
clayey sands, low plasticity
Soft sandy-silty clays to clayey
sands,
low to medium plasticity
Stiff to hard high plasticity clays
Very soft buy mud and silty
sands
Alluvium deposits, sandy-silty
clays to clayey sands-silts, low
strength and high
compressibility
Stiff to hard sandy-silty clays to
clayey sands
Very stiff to hard low to medium
plasticity clays to sandy
clays,overconsolidated with
rubbles and thin layers of
gravels
GreenSchists & Gneiss

VS
(m/s)
(5)

VP
(m/s)
(6)

200-350
(250)
300-400
(350)

4001700

8-20 (15)

1900

15-20 (20)

1800

20-25 (20)

1800

20-40 (30)

1800

20-25 (25)

150-250
(200)

1800

15-25 (20)

350-700
(600)

2000

6-30 (30)

700-850
(750)

3200

50-60 (60)

1750-2200
(2000)

4500

180-200
(200)

200-300
(250)
300-400
(350)
120-220
(180)

QS
(7)

Table 2. Damage distribution for the central part of the city: total number and percentage of buildings in each
damage state with and without mitigation for the scenario with Tm=475 years (*).

No
mitigation
Mitigation
option 1

Mitigation
option 2

Mitigation
option 3

Damage State

DS0

DS1

DS2

DS3

DS4

DS5

Total buildings

424

1988

1408

699

285

228

Percentage (%)

8.43

39.51

27.98

13.89

5.66

4.53

Total buildings

432

2034

1457

709

252

148

Percentage (%)

8.59

40.42

28.95

14.09

5.01

2.94

Change (%)

1.89

2.31

3.48

1.43

-11.58

-35.09

Total buildings

687

2264

1585

258

131

107

Percentage (%)

13.65

44.99

31.50

5.13

2.60

2.13

Change (%)

62.03

13.88

12.57

-63.09

-54.04

-53.07

Total buildings

486

2059

1442

671

238

136

Percentage (%)

9.66

40.92

28.66

13.33

4.73

2.70

Change (%)

14.62

3.57

2.41

-4.01

-16.49

-40.35

*Green: DS0 & DS1 (light mostly non-structural damage).


Yellow: DS2 & DS3 (repairable structural or non-structural damage).
Red: DS4 & DS5 (severe structural damage that would mostly be demolished).

393

Table 3. Mitigation options for R/C buildings in Thessaloniki.


Mitigation
option

Number &
percentage of
buildings in the
study area

262 (~5.2%)

3732 (~74.2%)

262 (~5.2%)

Building typology
without mitigation

Building typology after


mitigation

low-code, regularly infilled


frames

low-code, regularly infilled


dual systems

low-code, irregularly
frames (pilotis)

low-code, irregularly
infilled dual systems
(pilotis)

low-code, regularly infilled


frames

high-code, regularly
infilled frames

low-code, irregularly
frames (pilotis)

high-code, irregularly
frames (pilotis)

low-code, regularly infilled


dual systems

high-code, regularly
infilled dual systems

low-code, irregularly
infilled dual systems
(pilotis)

high-code, irregularly
infilled dual systems
(pilotis)

low-code, regularly infilled


frames

high-code, regularly
infilled dual systems

low-code, irregularly
frames (pilotis)

high-code, irregularly
infilled dual systems
(pilotis)

low-code, regularly infilled


frames

low-code, regularly infilled


dual systems

394

Seismic hazard scenarios


(100, 475, 1000 years)

Inventory data

Seismic response analysis


- site effects
(Microzonation study)

Definition of typology

Vulnerability models

Casualty model

Mitigation:
Definition of different
mitigation schemes

Loss estimation:
physical damage, casualties,
direct economic losses

Uncertainties
(due to inventory data,
seismic hazard, vulnerability
and casualty models)

Fig 1: Loss estimation modeling approach for Thessaloniki.

Fig 2: Map with the main geological units of the broader area of Thessaloniki city (IGME, 1978).

395

12

(a)

(b)

10

12

Main Road
Byzantine City Walls
0

1000

Main Road
Byzantine City Walls

2000

(c)

1000

2000

(d)

Main Road
Byzantine City Walls
0

1000

2000

Fig 3: Geotechnical thematic maps for the main soil formations (A, C, E and F and G). a) thickness of
formation A(artificial fills); b) thickness of formation C (sandy-silty & organic soils), c) top surface of stiff
clayey formations E-F; d) top surface of the formation G (bedrock).

396

Microzonation of
Thes s aloniki
A
B1
B2
B3
C

1.0

Microzonation of
Thes s aloniki

Based on Resonant Column Tests & Literature

Based on Resonant Column Tests & Literature

A
B1
B2
B3
C

D
E
F
G

0.8

D
E
F
G

20.0

G/Gmax

DS (%)

0.6

10.0

0.4

0.2

0.0
0.0001

0.001

0.01

0.1

10

0.0
0.0001

0.001

0.01

0.1

She a r Stra in, (%)

She a r Stra in, (%)

Fig 4: Average degradation curves of shear modulus G/Go- and damping ratio Ds- of the main soil
formations of Thessaloniki urban area.

Fig 5: Grid (500x500, 250x250m) of sites used for the 1D EQL theoretical analyses (n=520).

397

10

40.8

Tm=475
40.75
400
40.7

380
360

0.15g

340

40.65

320

0.20g
40.6

300
280

0.24g

260
240

40.55

0.28g

220
200

40.5

180
160
140

40.45

120
100

40.4
22.7

22.75

22.8

22.85

22.9

22.95

23

23.05

23.1

23.15

23.2

Fig 6: Representative results in terms of mean PGA(gs) at outcropping bedrock conditions for the area in
interest which, stemming from seismic hazard analysis for 475 years return period (Papaioannou 2004).

Fig 7: Distribution of mean peak ground acceleration (PGA: gs) obtained by 1D (EQL) analytical approach
(Tm=475 years).

398

Fig 8: Distribution of mean peak ground velocity (PGV: cm/s) obtained by 1D (EQL) analytical approach
(Tm=475 years).

Fig 9: Vulnerability assessment and damage distribution of Thessalonikis water system (Tm=475 years).

399

Fig 10: Distribution of damages to roadway bridges of Thessaloniki Metropolitan area (Tm=475 years).

Fig 11: Damage distribution for the central part of the city - predicted tagging of buildings (Tm=475 years)
(absolute values can be found in table 2).

400

Minimization of Fixed-Head Pile Bending at Optimal Radius under


Realistic Conditions by Using Seismic Deformation Method
M. Saitoh1
1

Saitama University, Japan

Abstract
In Saitoh (2005), the author described the fundamental relation between the radius and the
bending strains at the head of the pile using three-dimensional wave propagation theory,
showing the presence of an optimal pile radius that minimizes the bending strain at the head
of the pile. From a practical point of view, however, piles may be embedded not only in
homogeneous soil, but also in layered soil strata. Moreover, failure of the soil surrounding a
pile and slippage and separation at the interface between the pile and the soil may occur
during earthquakes. In recent years, a static numerical method called the Seismic
Deformation Method (SDM) has been used in practical applications as a useful method to
evaluate the response of a pile kinematically affected by the deformation of the soil stratum.
The SDM can easily consider realistic conditions of layered soil strata and the nonlinearity
of the soil in the vicinity of piles. The results of parametric analyses by the SDM show that
the relation between the radius (slenderness ratio) and the bending strain at the head of the
pile (normalized by the mean shear strain of a soil stratum) is close to the relation obtained
by a previously studied theoretical solution. Subsequently, the SDM is applied to numerical
studies performed for estimating the bending strain under practical conditions. Numerical
results show that the normalized bending strain due to the inertial bending tends to increase
for small slenderness ratio a H as the nonlinearity of the soil increases, whereas the
normalized bending strain due to the kinematic bending tends to decrease. Therefore, an
opposite change in the bending strains with a H , due to the inertial bending and the
kinematic bending, occurs due to the nonlinearity of the soil. As a result, the optimal radius
tends to increase due to the nonlinearity.

INTRODUCTION

This paper is one in a sequence of studies


[Saitoh (2005)] investigating the bending strains
at the head of a vertical, cylindrical fixed-head
pile and the effect of the pile radius on the
bending strains, in soil-pile-structure systems
where the kinematic interaction dominates.
Many studies have revealed that kinematic
bending induced by the lateral deformation of
soft soil surrounding a pile has become an
important problem in geotechnical engineering
[Mizuno, et al. (1984), Ohira et al. (1985),
Mizuno (1987), Tazoh et al. (1988), Kavvadas
and Gazetas (1993), Kaynia and Mahzooni
(1996), Mylonakis et al. (1997), Nikolaou et al.
(2001), Mylonakis (2001), and Luo et al. (2002)].

401

In Saitoh (2005), the author described the


fundamental relation between the radius and
the bending strains at the head of the pile using
three-dimensional wave propagation theory,
showing the presence of an optimal pile radius
that minimizes the bending strain at the head of
the pile. Moreover, the author derived criteria by
which the optimal radius can be determined for
soil-pile-structure systems. Saitoh (2005) gave
a general expression for the closed-form
formulae by normalizing the bending strain with
respect to a mean shear strain of a soil stratum
s . Consequently, it was found that the
normalized bending strains can be expressed
by normalized parameters, such as the
slenderness ratio a H , the ratio of soil and pile

stiffness E g E p , and a factor f r that


represents dynamic characteristics of loading at
the head of the pile and deformation of the soil.
From a practical point of view, however,
piles may be embedded not only in
homogeneous soil, but also in layered soil
strata. Moreover, failure of the soil surrounding
a pile and slippage and separation at the
interface between the pile and the soil may
occur during earthquakes. Therefore, the
characteristics of the optimal pile radius
evaluated based on three-dimensional wave
propagation theory in the previous study may
change due to such practical conditions.
However, closed-form formulae based on wave
propagation theory under such practical
conditions are extremely difficult to derive;
therefore, an alternative method for predicting
the effect of such practical conditions upon the
optimal pile radius is desirable.
In recent years, a static numerical method
called the Seismic Deformation Method (SDM)
has been used in practical applications as a
useful method to evaluate the response of a
pile kinematically affected by the deformation of
the soil stratum [Luo et al. (2002)]. There are
generally two types of methods for solving the
response of the pile by using SDM: 1) use of a
numerical solution based on discretized models;
and 2) use of an analytical solution based on
continuum models. The discretized SDM
method can easily consider realistic conditions
of layered soil strata and the nonlinearity of the
soil in the vicinity of piles. In addition, this
method can also consider the material
nonlinearity of piles. Luo et al. (2002) verified
the validity of the SDM in a simulative analysis
of pile foundations embedded in soft soil that
experienced the Hyogoken-Nanbu earthquake.
It is expected, therefore, that the effect of such
practical conditions upon the optimal pile radius
could be predicted by using the SDM. Until now,
however, it has not yet been verified whether
SDM is an appropriate method for estimating
the normalized bending strains and the optimal
pile radius, even for the simplest case of a
homogeneous elastic soil profile. Using the
continuum SDM method, a closed-form formula
of the normalized bending strains for the above
case can easily be obtained. Therefore, the
fundamentals of a kinematically affected pile,
such as the consistency of the normalized
parameters governing the normalized bending

402

strains, can directly be identified from the


closed-form formula. Therefore, this study
evaluates the adequacy of the SDM for
estimating the normalized bending strain and
the optimal radius. Subsequently, numerical
studies are performed for estimating the
bending strain under the nonlinearity of
cohesive soil as practical conditions.
The objectives of the present study are: (1)
To obtain a closed-form formula of the
normalized bending strains at the head of the
pile using SDM; (2) to compare the optimal pile
radius evaluated by the SDM with that
evaluated by the rigorous solution [Saitoh
(2005)]; and (4) to obtain numerical solutions of
the normalized bending strains at the head of a
pile in the presence of a local nonlinearity in
cohesive soil as realistic conditions by using the
SDM.
CLOSED-FORM FORMULA BASED ON SDM

The soil-pile-structure system is shown in


Fig. 1. A vertical, cylindrical pile of radius a is
embedded in a homogeneous elastic stratum of
thickness H . The toe of the pile is supported
by a compliant bedrock. Complete bonding at
the interface between the pile and soil medium
is assumed. The toe of the pile is presumed to
be restrained elastically against rotational
movements by a spring of static stiffness K r at
the base. In this study, the stiffness K r
assumed is described by the following formula
[Borowicka (1943)]:
Kr =

8a 3 bVsb2
,
3(1 b )

(1)

where b , V sb , and b are the density, the


shear velocity, and the Poissons ratio of the
compliant bedrock, respectively.
Referring to a previous study [Luo et al.
(2002)], the equilibrium equation of the pile is:

EpI

4u p
z 4

= Kd (u p u s )

(2)

where E p , I , and d (= 2a ) are the Young's


modulus, geometrical moment of inertia, and
the diameter (radius) of the pile, respectively;

is the horizontal deformation of the


subgrade; and u p and u s are the horizontal

Kd = 1.2 E g

deformation of the pile and the soil, respectively.


The product Kd can be approximated by
[Gazetas and Dobry (1984)]:

Fixed-Head Conditions

us

2a

(3)

Rotational Stiffness at the


Toe of the pile

where E g is the Young's modulus of the soil.


The deformation of the soil u s is described
in one-dimensional wave propagation theory as

u s = u suf sin

Vs

1
0

b p

a p

where u suf is the maximum value of the


relative displacement of the surface ground with
respect to the bedrock; is the fundamental
frequency of the soil stratum; and Vs is the
shear wave velocity of the soil.
Solving Eq. 1 for the bending strain at the
head of the pile normalized by a mean shear
strain of the soil medium yields
4

p
s
e

1 + 0 l r s r1
4

(C

p H

p H

[e (A
pH

1
1 2 l r s
4

4
r

(6)
ap = e

pH

where

Eg 4
Kd
a
, sr =
, p =
4E p I
H
Ep

2
4

2.467 , and 2 =

bp = e

p H

bm = e

p H

Eg
Eb

1
0

b p

a p

5 5
V

=
3.985 ,
, 0
fr =
384
s EpH 2
1

1 =

(cos H + sin H ) ,
(cos H + sin H ).
(cos H sin H ) ,
(cos H sin H ),

p H

, and g r =

3
3 1 b2

rr s r 4 ,
2 2

where E b is the Young's modulus of the


compliant bedrock.
The coefficients AI , B I , C I , and DI are
determined by the following equations:

(5)

lr =

p H

am = e

rr =

[ (A sin H B cos H )
(C sin H D cos H )]
I

1
0 AK 0
1 1 + g r B K 1
= ,
bm C K 0
am

bm a m D K 0

0
1 gr
ap
bp

where

sin p H B K cos p H )

sin p H D K cos p H )

+ f r s r 4 l r 2 2 e
e

1
4
r

Kr

Fig. 1: Analytical soil-pile-structure model for a


homogeneous soil medium and a fixed-head pile
supported by rotationally compliant bedrock.

(4)

4 4 5
1.620 .
6

0
1 gr
ap
bp

1
0 AI 0
1 1 + g r B I 0
= .
bm C I 0
am

bm a m D I 1
(7)

The coefficients AK , B K , C K , and DK are


determined by the following equations:

The first and second terms of Eq. 5 are


associated with the kinematic bending and the
inertial bending, respectively. It is noted that
the coefficient of the second term, f r ,

403

Fixed-Head Conditions

expresses the effect of the lateral load relative


to the deformation of the soil medium. This
coefficient also appears in the closed-form
formula of the theoretical solution [Saitoh
(2005)]. In that previous study, this factor f r
was treated as a complex value since a phase
lag generally appears between the lateral load
V and the mean shear strain of the soil
medium s . Therefore, this factor is written as:

f r = Fr e ir ,

h1
h2

h3

hi

Layer-2

usi

Layer-i

#
#

hN

Layer-1

Layer-3

us1
us 2
us 3

usN

Layer-N

Rotational Stiffness at the

(8)

Toe of the pile

where

Fr =

Kr

Fig. 2: Discretized soil-pile-structure model for


layered soil medium and a fixed-head pile
supported by rotationally compliant bedrock.

V
and
s EpH 2
1

V
.
s

r = arg

pile is discretized by beam elements; and (2)


the subgrade spring of the soil is concentrated
at each node of the beam elements. Providing
that the soil and the pile are discretized with a
finite number of layers N, the horizontal
subgrade spring of the soil k i at the i-th node of
the pile is determined based on the following
formulae:

Here, the factor Fr is the ratio of the


maximum values of the lateral load V and the
mean shear strain s , whereas the factor r is
the phase lag of the lateral load V with respect
to the mean shear strain s .
Eq. 5 shows that the normalized bending
strain consists of the following normalized
parameters: (1) the slenderness ratio a H ; (2)
the ratio of the stiffness of the soil and the pile
E g E p ; (3) the ratio of the stiffness of the soil

K 1e =
K ie =

and the complaint bedrock E g Eb ; and (4) the


factors Fr and r . It is noted, therefore, that
the normalized parameters that govern the
normalized bending strain evaluated by the
SDM are consistent with those evaluated based
on the three-dimensional wave propagation
theory.

k1
h1 ( i = 1 )
2

(9a)

k i 1 hi 1 k i hi
( i > 1 ), (9b)
+
2
2

where k i is the product of the diameter of


the pile d and the horizontal deformation of the
subgrade of the i-th layer K i ; and hi is the
thickness of the i-th layer. According to Gazetas
and Dobry (1984), the product k i ( = K i d ) in Eq.
9 can be approximated by

k i = 1.2 E gi ,

NUMERICAL SOLUTION BASED ON


DISCRETIZED SDM

(10)

where E gi is the Young's modulus of the

As described above, a numerical method


based on the discretized SDM is convenient for
practical applications. The soil-pile-structure
system used is shown in Fig. 2. The governing
equation of this system is analogous to that of
the system used for the analytical solution
shown in Eq. 1. Differences between the
systems can be summarized as follows: (1) the

soil in the i-th layer.


The numerical procedures used in this study
are briefly explained below. The phase lag r
can be directly incorporated into the analytical
solution of SDM as well as the theoretical
solution [Saitoh (2005)] since these solutions

404

(b)

(a)
0.4

0.4

0.3

0.3
0.2

p/s

0.1

0.1

-3
-4

0.0

0.1

0.0

0.3

-6

0.4

-6

0.4

0.3

p/ S

p/S

0.3

(d)

0.4

0.3

0.2

0.2

0.1

0.1

0.0
0.0

0.1

0.2

0.3

a/H

0.0
0.0

0.4

0.1

0.2

a/H

0.3

0.4

(f)

(e)
0.4

0.4

0.3

0.3

0.1

-3
-4

0.0

0.1

-5

0.2

a/H

0.3
0.4

-6

0.1

-3
-4

0.0

log
(F
r)

0.0

0.2

0.0

0.1

-5

0.2

a/H

log
(F
r)

0.2

p/s

p/s

-5

0.2

a/H

(c)

0.4

0.1

-5

0.2

a/H

-4

0.0

log
(F
r)

0.0

-3

log
(F
r)

p/ s

0.2

0.3
0.4

-6

Fig. 3: Variation of normalized bending strains with based on numerical SDM ( N = 20)
[(a) inertial bending, (c) kinematic bending, and (e) total bending]. Comparison with the
theoretical solution of Saitoh (2005) [(b), (d), and (f)].

can mathematically be treated as complex


values. On the other hand, the numerical
solution of SDM is generally evaluated by an
incremental static analysis; therefore, the effect
of the phase lag r should be considered using
an appropriate technique. In this study, a

sinusoidal wave form is assumed in each of the


lateral load V and the displacements of the soil
medium associated with the mean shear
strain s . The phase lag r is given to the
with respect to the
lateral load V
displacements of the soil medium. According to

405

a well-known step-by-step procedure, the waveformed lateral load V with the phase lag r is
given to the pile head, whereas the waveformed displacements of the soil medium u si
are simultaneously given to the nodes of the
discretized beam elements of the pile, as
shown in Fig. 2.

the inertial bending and the kinematic bending,


a local minimum area appears where the
bending strains due to both the inertial bending
and the kinematic bending are minimized, as
shown in Fig. 3(e). The pile radius related to
this slenderness ratio a H minimizing the
normalized bending strains is defined as an
optimal pile radius in this study. This behavior
was also observed in the theoretical solution
shown in Fig. 3(f). It was found that the
normalized bending strains evaluated by the
SDM are quite similar to those evaluated by the
theoretical solution.

NORMALIZED BENDING STRAIN AND OPTIMAL


PILE RADIUS

Comparisons with numerical SDM and


theoretical solutions
Fig. 3
bending
numerical
functions

APPLICATION TO REALISTIC CONDITIONS

shows the variations in normalized


strains evaluated by using the
solutions of the SDM ( N =20), as
of the slenderness ratio a H with

This section shows the application of the


SDM to practical conditions. In this study, the
effect of local nonlinearity in soil is considered.
It is assumed that the soil spring has a bilinear
skeleton curve where the ratio of the second
stiffness (after yielding) to the initial stiffness is
assumed to be zero; this is identical to the
subgrade reaction that follows the elastic /
perfectly plastic constitutive law. The ultimate
soil reaction depends on the type of soil
supporting the pile (Broms (1964), Matlock
(1970), Reese (1975)). In this study, it is
assumed that the pile is embedded in a
cohesive soil. In general, the following
expression for the ultimate soil reaction per unit
length of a pile embedded in a cohesive soil is
used:
p i = S u d ,
(11)

different values of the factor Fr for the phase

lag r = 0 . The absolute values of the


normalized bending strains and the related
terms are shown. It is assumed in Fig. 3 that
the stiffness ratio E g E p = 0.001 ; the stiffness
ratio E g Eb = 0.05 ; and the Poisson's ratio

b = 0.45 . Fig. 3(a) indicates that the


normalized bending strain due to the inertial
bending rapidly decreases as a H increases.
When the factor Fr increases, the strain
gradually increases within the range of a H
shown. This behavior is similar to that found by
the theoretical solution [Saitoh (2005)]. The
corresponding published results of the
theoretical
solutions
are
simultaneously
presented in Fig. 3, where it was assumed that
the mass density ratio p g = 1.25 ; the

where = 9 is used for soft clay, and S u is


the soil undrained shear strength. In this study,
S u is assumed to be 30 kN/m2. Matlock (1970)
proposed another expression for shallow depth.
Therefore, both of the expressions are used
according to the depth in this study.
Fig. 4 shows the variations in normalized
bending strains as functions of the slenderness
ratio a H with different values of the factor Fr

material damping hg = 0.05 ; and the Poissons


ratios = b = 0.45 . The other normalized
parameters were identical to those used in the
SDM. Fig. 3(c) shows the normalized bending
strain due to the kinematic bending. A single
curve is shown in Figs. 3 (c) and (d) because
this bending strain is independent of the factor
Fr . The appearance of the local maximum and
the almost-linear increase up to the maximum
are similar to the theoretical solution shown in
Fig. 3(d). With an opposite change in the
normalized bending strains with a H , due to

for the phase lag r = 0 in the case of

s = 0.01 (i.e., changes in the lateral load V

are accompanied by changes in the factor Fr ).


Assumptions identical to those in Fig. 3 are
made in Fig. 4 for the non-dimensional
parameters E g E p , E g Eb , and p g , and
the Poissons ratios and b . Herein, Figs. 4

406

(b)

(a)

0.4
0.4

0.3

0.2

p/S

0.1

-4
0.1

0.3
0.4

-6

(c)

0.2

a/H

0.3

0.4

(d)

0.4

20
15

H non (m)

0.3

p/s

0.1

-5

0.2

a/H

0.0
0.0

log
(F
r)

0.0

0.1

-3

0.0

0.2

0.2

10

0.1

-4
0.1

-5

0.2

a/H

0.3
0.4

-6

-3
-4

0.0

log
(F
r)

0.0

0.0

-3

0.1

-5

0.2

a/H

log
(F
r)

p/ s

0.3

0.3
0.4

-6

s = 0.01 for (a) inertial


bending, (b) kinematic bending, (c) total bending, and (d) inelastic region H non . ( r = 0 ,
E g E p = 0.001 , E p = 2.5 10 7 kN/m2 and p = 2.5 t/m3.)

Fig. 4: Variation of normalized bending strains with

Fr

in case of

strain shifts toward smaller values due to the


nonlinearity of the soil. It is noted, therefore,
that an opposite change in the bending strains
with a H , due to the inertial bending and the
kinematic bending, occurs due to the
nonlinearity of the soil.
Figs. 3(c) and 4(c) indicate that the distance
between the slenderness ratios at the local
minimum and the local maximum tends to
decrease due to the nonlinearity of the soil. In
addition, the amplitudes of the normalized
bending strains at the local minimum and the
local maximum tend to be close because of the
opposite change in the bending strains due to

(d) shows the maximum length from the head of


the pile to the soil spring that exceeds the
ultimate strength of the soil, when the pile is
influenced by both the inertial bending and the
kinematic bending. Figs. 3(a) and 4(a) indicate
that the normalized bending strain due to the
inertial bending tends to increase for small a H
( 0.15) due to the nonlinearity of the soil.
On the other hand, Fig. 3(b) and 4(b) show
that the local maximum of the bending strain
due to the kinematic bending decreases above
a H = 0.15 and the corresponding slenderness
ratio that maximizes the kinematic bending

407

Foundations (ed. Nogami, T.), pp. 53-78, New


York: ASCE.
Tazoh, T., Wakahara, T., Shimizu, K., Matsuzaki, M.
(1988) "Effective motion of group pile
foundations". Proc., Proceedings of the 8th
World Conference on Earthquake Engineering,
Vol.3, Tokyo, 587-592.
Kavvadas, M., Gazetas, G. (1993) "Kinematic
seismic response and bending of free-head
piles in layered soil". Geotechnique, 43(2), 207222.
Kaynia, A. M., Mahzooni, S. (1996) "Forces in pile
foundations under seismic loading". Journal of
Engineering Mechanics, ASCE, 122(1), 46-53.
Mylonakis, G., Nikolaou, A.S., Gazetas, G. (1997)
"Soil-pile-bridge seismic interaction: kinematic
and inertial effects. Part I: soft soil". Earthquake
Engineering and Structural Dynamics, 26, 337359.
Nikolaou, S., Mylonakis, G., Gazetas, G., Tazoh, T.
(2001) "Kinematic pile bending during
earthquakes: analysis and field measurements".
Geotechnique, 51, 425-440.
Mylonakis, G. (2001) "Simplified model for seismic
pile bending at soil layer interfaces". Soils and
Foundations, JGS, 41(4), 47-58.
Luo, X., Murono, Y., Nishimura, A. (2002) "Verifying
adequacy of the seismic deformation method by
using real examples of earthquake damage".
Soil Dynamics and Earthquake Engineering, 22,
17-28.
Borowicka, H. (1943) "ber ausmittig belastete
starre
Platten
auf
elastischisotropem
Untergrund". Ingenieur-Auchiv, 1:1-8.
Gazetas, G., Dobry, R. (1984) "Horizontal response
of piles in layered soils". Journal of
Geotechnical Engineering, ASCE, 110(6), 937956.
Broms, B. B. (1964) "Lateral resistance of piles in
cohesionless soils". Journal of Soil Mechanics
and Foundation Division, ASCE, 90(SM3), 123156.
Broms, B. B. (1964) "Lateral resistance of piles in
cohesive soils". Journal of Soil Mechanics and
Foundation Division, ASCE, 90(SM2), 27-63.
Matlock, H. (1970) "Correlations for design of
laterally loaded piles in soft clay". Paper No.
OTC 1204, Proc. 2nd Annual Offshore
Technology Conference, Houston, Texas, I, 557594.
Reese, L. C., Cox, W. R., and Koop, F. D. (1975)
"Field testing and analysis of laterally loaded
piles in stiff clay". Paper No. OTC 2312, Proc.
7th Annual Offshore Technology Conference,
Houston, Texas, II, 672-690.

the inertial bending and the kinematic bending.


The opposite change also induces a slight
change in the slenderness ratio associated with
the local minimum area: in fact, the local
minimum area tends to shift toward larger
slenderness ratios.
CONCLUSIONS

1) This study verifies the adequacy of the


seismic deformation method (SDM) for
estimating an optimal pile radius that minimizes
the bending strains of fixed-head piles
embedded in a homogeneous soil in soil-pilestructure systems where the kinematic
interaction dominates.
2) Parametric studies show that the
variations in normalized bending strains with
a H evaluated from the numerical solution of
the SDM are quite close to those evaluated
from the theoretical solution.
3) The SDM is applied to numerical studies
for estimating the bending strain under the
nonlinearity in cohesive soil as realistic
conditions. More studies are necessary for
revealing the general effect of the nonlinearity
on the bending strain and the optimal radius.
REFERENCES
Saitoh, M. (2005) "Fixed-head pile bending by
kinematic interaction and criteria for its
minimization at optimal pile radius". Journal of
Geotechnical
and
Geoenvironmental
Engineering, ASCE, 131(10): 1243-1251.
Saitoh M. (2005) "Minimization of Fixed-Head Pile
Bending at Optimal Pile Radius", Proceedings
of the 1st Greece-Japan Workshop(Athens),
Seismic Design, Observation, Retrofit of
Foundations, pp.6167.
Mizuno, H., Iiba, M., Kitagawa, Y. (1984) "Shaking
table testing of seismic building-pile-twolayered- soil interaction". Proceedings of the 8th
World Conference on Earthquake Engineering,
Vol. 3, Earthquake Engrg. Res. Inst. (EERI).
San Francisco, California 649-656.
Ohira, A., Tazoh, T., Nakahi, S., Shimizu, K. (1985)
"Observation and analysis of earthquake
response behavior of foundation piles in soft soil
deposit". Journal of Structural Mechanics and
Earthquake Engineering, JSCE, 362(I-4), 417426, in Japanese.
Mizuno, H. (1987) "Pile damage during earthquakes
in Japan". Dynamic response of Pile

408

Seismic Response Analysis of Pile Foundation


using Finite Element Method
T. MAKI1, S. TSUCHIYA2 , T. WATANABE3 , K. MAEKAWA4
1

Saitama University, Japan


COMS Engineering Co, Ltd., Japan
3
Hokubu Consultant Co, Ltd., Japan
4
The University of Tokyo, Japan

Abstract
In this paper, 3-D and 2-D seismic response analyses of RC pile-soil system were
conducted using nonlinear finite element method, and the effect of the thickness of soil
elements used in the 2-D models on the response of pile foundation was investigated by
comparing these analytical results. In the results obtained by the 2-D models with different
values of soil element thickness, macroscopic response at the top support was similar,
whereas the damage level in the RC piles was remarkably different. For the target structure
in this paper, the 2-D model in which the thickness of soil elements was equal to the footing
width could give an equivalent result to the 3-D model.

INTRODUCTION

After the 1995 Hyogoken Nanbu Earthquake,


the seismic performance verification method
has been improved. The seismic response
analysis of RC pile foundation-soil coupled
system has been applied to the damage
assessment and seismic diagnostics of RC
structures. In the JSCE Standard Specifications
for Concrete Structures (JSCE 2005), the
seismic response analysis using a finite
element method (FEM) of soil-structure coupled
system with input seismic excitation at the
engineering base layer is already prescribed as
a principle to obtain the seismic structural
response.
However, there are still rare cases where a
response analysis has been applied using 3-D
FEM based on the nonlinear constitutive laws of
reinforced concrete and soil for the seismic
performance verification in the practical design
of RC structures. Concerning the seismic
response of such coupled system, 2-D analysis
is still in use as an alternative method to 3-D
analysis. In the 2-D FEM, structures and soil
are usually modeled with plane stress and
plane strain elements, respectively. However,
the obtained response by the 2-D analysis is
highly influenced by the model such as

409

thickness of soil element (Ishihara 1994),


especially for the pile foundation of which
response contains so-called '3-D effect'.
Therefore, when such an alternative method to
3-D analysis is applied, the careful engineering
judgment
in
the
modeling
and
the
understanding of the obtained results are
needed.
The 3-D finite element analyses of structurepile foundation-soil coupled system have been
conducted by some researchers (Kimura and
Zhang 2000). However, there has never been
such an analysis that is performed using 3-D
FEM which can consider not only flexural
nonlinearity but also shear failure and postpeak behavior of RC members and structures,
together with soil nonlinearity.
In this paper, the seismic response analysis
of the abutment supported by pile foundation
was conducted using 3-D FEM. In the analysis,
both structure and soil were modeled by 3-D
solid elements with nonlinear material
constitutive laws. In addition, keeping in mind
the application of such coupled response
analysis and subsequent feedback to practical
design, 2-D seismic response analysis was also
performed and the result was fully compared
with that from the 3-D analysis in terms of the
sensitivity of the thickness of soil elements.

TARGET STRUCTURE

in the analysis. This code has already been


verified for the static and dynamic behavior of
various types of RC structures. 3-D finite
element mesh is shown in Fig 3. Both structure
and soil were modeled by 20-node
isoparametric 3-D solid elements. The soil up to
the depth of G.L. -18.4m was included in the
model and the one-half of the structure by the
plane along with the bridge axis was modeled.
The circular cross section of RC piles was
substituted by the equivalent square section
having the equal moment of inertia. As shown
in Fig 3(b), 16-node viscous boundary elements
produced by reducing the degrees of freedom
in 20-node solid elements were provided
around the surrounding soil of the foundation.
Moreover, 16-node joint interface elements
were provided between soil and piles or footing,
in order to consider contact, separation and slip.
The resultant model consisted of 24,290 nodes
and 5,762 elements.
Fig 4 shows the 2-D finite element mesh.
The abutment and pile foundation were
modeled by 8-node plane stress plate element,
and the soil was modeled by 8-node plane
strain plate element. The mesh division in the 2D model had perfect consistency with that in the
3-D model. 6-node viscous boundary elements

Fig 1 and Fig 2 show the detail of the target


structure in the response analysis and the soil
profiles at the site, respectively. It was the
abutment of 4-span continuous girder bridge
supported by cast-in-place RC pile foundation.
The diameter of the pile was 1.2m and the
deformed bars having 32mm diameter were
arranged as longitudinal reinforcement. The
vertical design load at the shoe was 2,700kN.
This is the actually-existing structure; however,
the surrounding soil was improved over 4.5m
width around the piles and 6.0m depth under
the footing, and the seismic performance of the
structural system was verified in its design
process. In this paper, the effect of soil
improvement is not considered in order to focus
on the nonlinear plastic response of the piles,
as well as to exclude the uncertainty in the
mechanical behavior of improved soil.
NUMERICAL METHOD
3-D and 2-D Finite Element Models

The 3-D nonlinear finite element analysis


code for RC structures, named "COM3"
(Okamura 1991, Maekawa 2003), was applied
B-B

3,640

3,124

2,000
A

4@3,025=12,100

Cast-in-place Pile
1,200
L=13.0m
n=15
2,000

12,148

10,948

14,500

600

1,176

1,200

8,500

A-A

0.00

B
8,000

1,200

1,176

600

As

2.65
3.70

600 1,400 2,056

4,582

40

50

3
12

Ags
4

As

600

N-SPT
20 30

10
3

5.00

12,900

3,124

C-C

D-D

Ac1-3

Ac1-2
9.95

1,200

1
1/35

8.00

3,640

Ac1-1

1/35
3

Ag

300 2,000

12.50

2432

1,200

2@3,050=6,100

1,200

Cast-in-place
Pile
1,200
L=13.0m
n=15

Detail of Pile

Nsl
16.00

Fig 2: Soil Profiles

410

42
50/10

Unit: mm

Fig 1: Detail of Target Structure

30

50/8

160

6,765

50/10
50/8

18.4m

25.665m

(a)

157.3 m

.65
44

(b)

Fig 3: (a) 3-D Finite Element Mesh, and (b) Viscous Boundary Elements
Viscous Boundary Elements

Fig 4: 2-D Finite Element Mesh

produced by reducing the degrees of freedom


in 8-node plate elements were provided, as well
as the 6-node joint interface elements were
provided between soil and piles or footing. As
already mentioned in the first section, the
thickness of soil elements in the 2-D model
should be carefully determined in terms of the
3-D effect in the response of the pile foundation.
In order to discuss the influence of the soil
thickness, the three types of the model (Type A,
B and C) were investigated, which had the
thicknesses equal to the footing width (Type A:
7.25m), 1.5 times as large as the footing width
(Type B: 11.65m) and 3.0 times as large as the
footing width (Type C: 20.65m), respectively.
Material Constitutive Laws and Parameters

The nonlinear path-dependent constitutive


law for reinforced concrete was applied to the
3-D solid and the 2-D plate elements for RC
piles (Okamura 1991, Maekawa 2003). This

411

constitutive law based on the smeared crack


concept, including the non-orthogonal multidirectional fixed crack model for concrete and
the buckling model for reinforcing bar, has
already been verified for various types of RC
structures. Moreover, the nonlinear shear
deformation after cracking can also be precisely
evaluated in this constitutive law. The
parameters for the above-mentioned RC solid
and RC plate elements were the compressive
and tensile strength of concrete (f'c = 24N/mm2
and ft = 1.914N/mm2, respectively) and the yield
stress of reinforcing bar (fy = 345 N/mm2).
In the nonlinear constitutive law for soil, the
Ohsaki's model (Ohsaki 1980) was applied to
the deviatoric (shear) stress-strain relationship.
The initial shear modulus G0 and the shear
strength Su were determined by the following
equations (1) and (2), according to the soil
profiles and N-SPT values:

Acceleration (gal)

900

Scale: X10

JSCE Level2 Inland-type1

600
300
0

Gap

-300
-600

Max: 749.65
Min: -604.73

Analysis Target (12sec)

-900
0

10
Time (sec)

15

Soft Silt Layer

20

Gap

Fig 5: Input Seismic Acceleration Wave (Original)


G0 = 11.76 N 0.8 [N/mm2]

(1)

Fig 6: Deformation of Pile Foundation (3.39sec)

Su = G0 / 1100 [N/mm ] (for Sand)

0.3

(2)

Response Displ. (m)

= G0 / 600 [N/mm2] (for Clay)


where, N : N-SPT value.

The volumetric component in the soil model


was assumed to be linear elastic. The bulk
modulus was obtained from the initial shear
modulus and the Poisson's ratio (assumed as
0.3). The dilatancy due to cyclic shearing, as
well as the confining stress dependency in the
shear strength, was ignored.
The normal and tangential viscous damping
coefficients for the boundary element were
determined by the equations from (3) to (5):
n = VP =
s = VS

2(1-)
VS (Normal)
1-2

(3)

(Tangential)

(4)

Vs = G0 /
where,

VP
VS

3D
Max/Min

0.2
0.1
0
-0.1

Max: 0.2102
Min: -0.2245

-0.2
-0.3

Response Accel. (gal)

900

10

12

600
300
0
-300
3D
Max/Min

-600

Max: 1628.0
Min: -757.10

-900
0

6
Time (sec)

10

12

Fig 7: Response Displacement and Acceleration


at the Top Support

as the wave used for seismic verification, and


was composed by the response spectra on the
fault using attenuation model and the phase
characteristics according to the asperity and
rupture process of the fault, based on many
observed records. In the analysis, the maximum
acceleration level was parametrically changed
as 2/3 and 1/3 of the original value. In this
paper, these waves are called "2/3 Wave" and
"1/3 Wave", respectively.

(5)

: Density of Soil,
: Poisson's Ratio,
: Primary Wave Velocity, and
: Shear Wave Velocity.

No other damping in the system except for


material hysteretic damping was considered.
The joint interface element between soil and
structure had minute normal tensile and shear
rigidities, and had high compressive rigidity to
avoid numerically the overlapping of soil and
RC elements.

RESULTS OF 3-D RESPONSE ANALYSIS


Macroscopic Response Behavior

Fig 6 shows the deformation of the pile


foundation when the maximum lateral response
displacement at the top support was observed
(3.39sec). Here, the lateral displacement at the
support was defined as the relative one to the
bottom end of the center pile. According to the
figure, the deformation of the foundation was
mainly caused by that of soil in the right-hand
side of the foundation. The large gap along the
left pile from the footing to the soft silt layer was
observed. In addition, at around the boundary

Boundary Conditions and Input Acceleration

All nodes at the bottom surface of the model


were perfectly restrained, and the nodes in the
free field soil elements outside the viscous
boundary were confined in the vertical direction.
Fig 5 shows the input acceleration wave.
This was the Level-2 Inland-type outcrop (2E)
wave prescribed in the JSCE code (JSCE 2005)

412

of this soft silt layer and the stiff sand layer


below, the gap was also observed in the righthand side of the pile.
Fig 7 shows the time histories of the
response displacement and acceleration at the
top support. The large response was observed
in both positive and negative sides during the
initial 4 secant, followed by the one-side
response in the negative side. The large values
of acceleration response at 4 and 10sec were
caused by the collision of soil and abutment.

Plane A

Plane A
Depth from G.L. (m)

-3

Negative
Disp.

-6

Positive
Disp.

Negative
Disp.

Left
Center
Right

-9

As
Positive
Disp.

Left
Center
Right

Ac

Ag
-12

Ag
Nsl

-15

-90

-60

-30

30

Curvature (10-6/cm)

Deformation of Foundation and Surrounding Soil

(a) Original Wave

Fig 8 shows the deformation of the pile


foundation and the contour plot of the vertical
strain of the pile when the maximum response

60

-30

-20

-10

10

(b) 1/3 Wave

Fig 9: Curvature Distributions of Piles (Plane A)


at Maximum/Minimum Displacement

As
Ag

Ac

Ag
Ag
Nsl

Overall Deformation

20

Curvature (10-6/cm)

Scale: x10

Plane A =>
Plane B =>
Plane C =>

Ag

Plane A

Plane B

Plane C

Fig 8: Deformation of Pile Foundation and Contour Plot of Vertical Strains (Original Wave)
Scale: x10

Plane
PA

Plane
QB

Plane
RC

Plane
SD

Q
R
S

(a) Original Wave (3.39sec)


Scale: x30

Plane
AP

Plane
BQ

Plane
R
C

Plane
DS

Q
R
S

(b) 1/3 Wave (1.91sec)


Fig 10: Deformation and Normal Strain Contour of Soil around Foundation

413

RESULTS OF 2-D RESPONSE ANAYSIS AND


COMPARISON WITH 3-D ANALYSIS

displacement at the top support was observed


(3.39sec), and Fig 9 shows the curvature
distributions of the piles in Plane A in the cases
with the original input wave and the moderate
wave. Strains at pile-footing connections and at
the boundary between the soft silt layer and the
stiff sand layer below were relatively large.
Slight difference in strains and curvatures could
be observed between the locations of piles in
and perpendicular to the direction of excitation.
Fig 10 shows the soil deformation and the
contour plot of the normal strains in the
horizontal direction when the maximum
response displacement at the top support was
observed. High compressive strains in the
horizontal direction could be observed in the
right-hand side of the foundation, i.e., in the soil
behind the foundation in the major response
direction (from Plane Q to S). This indicated
that the foundation deformed due to being
pushed by the soil at the rear side. The soil in
the left-hand side preceded the foundation in
the horizontal displacement, resulting in the
tensile strain localization of soil between the
front piles (left piles). In addition, according to
the compressive strain distribution, the effective
width of soil to the response of foundation was
found to be close to the width of footing.

The detailed response behavior of pile


foundation was discussed in the previous
section based on the results of 3-D finite
element analysis. In this section, the effect of
the thickness of soil elements on the response
of pile foundation was investigated according to
the comparison of the analytical results
obtained by three types of 2-D finite element
mesh with different thickness of soil elements.
Macroscopic Response Behavior

Fig 11 shows the overall deformation of the


soil-pile foundation system using Type A and C
meshes. Although the soil deformation in the
left-hand side of the foundation found to be
slightly large compared with the 3-D analysis
result shown in Fig 6, overall deformation
patterns in the both 3-D and 2-D analyses were
similar.
Fig 12 shows the time histories of response
displacement and acceleration at the top
support obtained by the three types of 2-D
meshes. No remarkable difference due to the
soil thickness could be seen for both original
and moderate input acceleration waves.

Scale: x10

Scale: x10

(a) Type A (3.37sec)

(b) Type C (3.39sec)

Fig 11: Overall Deformation in 2-D Analysis (Original Wave)


0.09
2D-A
2D-B
2D-C

0.2
0.1

Response Displ. (m)

Response Displ. (m)

0.3

0
-0.1
-0.2

0.03
0
-0.03
-0.06
-0.09
600

Response Accel. (gal)

-0.3
900

Response Accel. (gal)

2D-A
2D-B
2D-C

0.06

2D-A
2D-B
2D-C

600
300
0
-300
-600
-900

2D-A
2D-B
2D-C

400
200
0
-200
-400
-600

6
Time (sec)

10

12

(a) Original Wave

6
Time (sec)

(b) 1/3 Wave

Fig 12: Response Displacement and Acceleration at the Top Support in 2-D Analysis

414

10

12

Table 1: Maximum / Minimum Response at Top Support


3D

2D-A

Displacement

Wave

Acceleration

Displacement

2D-C

2D-B
Acceleration

Displacement

Acceleration

Displacement

Acceleration

Value, m Time, sec Value, gal Time, sec Value, m Time, sec Value, gal Time, sec Value, m Time, sec Value, gal Time, sec Value, m Time, sec Value, gal Time, sec

1.60
3.39
1.56
3.37
1.50
1.91

1628.0
-757.1
1147.0
-704.1
548.1
-485.0

9.88
4.10
9.83
4.06
9.78
3.94

0.2162
-0.2206
0.1118
-0.1402
0.0371
-0.0717

1.59
3.37
1.55
2.04
1.49
1.89

716.8
-521.6
493.7
-371.5
326.8
-283.1

8.58
8.36
8.54
4.43
1.81
4.39

0.2179
-0.2168
0.1115
-0.1464
0.0368
-0.0741

1.60
3.38
1.56
3.36
1.50
1.91

661.6
-501.9
448.4
-352.2
301.4
-231.9

8.56
8.37
8.55
8.35
1.80
4.38

0.2155
-0.2099
0.1091
-0.1490
0.0357
-0.0719

1.61
3.39
1.57
3.37
1.51
1.93

577.1
-455.2
427.3
-366.5
278.0
-197.2

As
Ag

Scale: x10
Ac

Ag
Ag
Nsl

(a) Type A (3.37sec)

(b) Type B (3.38sec)

(c) Type C (3.39sec)

Fig 13: Deformation of Pile Foundation and Contour Plot of Vertical Strains (Original Wave)
0

2D-A
2D-B

2D-C
3D

2D-A
2D-B

2D-C
3D

-6

-9

-12

2D-A
2D-B

2D-C
3D

As

-3
Depth from G.L. (m)

-3
Depth from G.L. (m)

-3

-6

-9

-12

Ag

-6

Ac
-9

Ag
-12

Ag
Nsl

-15
-120 -80

-40

40

80

-15
-120 -80

120

Curvature (10-6/cm)

-40

40

80

-15
-120 -80

120

Curvature (10-6/cm)

-40

40

80

120

Curvature (10-6/cm)

(a) Original Wave


0

2D-A
2D-B

2D-C
3D

2D-A
2D-B

2D-C
3D

-3

-6

-9

-12

2D-A
2D-B

2D-C
3D

As

-3
Depth from G.L. (m)

-3
Depth from G.L. (m)

1/3

0.2102
-0.2245
0.1064
-0.1563
0.0330
-0.0748

Depth from G.L. (m)

2/3

Max
Min
Max
Min
Max
Min

Depth from G.L. (m)

Original

-6

-9

-12

Ag

-6

Ac
-9

Ag
-12

Ag
Nsl

-15
-30

-20

-10

10

Curvature (10-6/cm)

20

30

-15
-30

-20

-10

10

Curvature (10-6/cm)

20

30

-15
-30

-20

-10

10

20

30

Curvature (10-6/cm)

(b) 1/3 Wave


Fig 14: Curvature Distributions of Piles at Maximum / Minimum Response Displacement

415

8.57
8.39
1.90
8.36
1.82
9.59

CONCLUSIONS

Moreover, these histories were almost similar to


those in the 3-D analysis shown in Fig 7, except
that the pulse-shaped acceleration response
due to the collision of soil and abutment
disappeared. The maximum and minimum
values of response displacement and
acceleration for all the cases were tabulated in
Table 1. There could be observed no difference
in the time between all the 2-D cases when the
maximum and minimum response occurred,
even in the comparison with the 3-D analysis
result. From the viewpoint of macroscopic
response in the soil-pile system, no remarkable
effect of the modeling could be found in these
results, and whichever 2-D model could give
similar results to the 3-D analysis.
Deformation of Pile Foundation

Fig 13 shows the deformations and the


contour plots of axial strains in the piles. In
terms of the flexural damage of the piles, slight
difference could be observed at the boundary of
the soft silt layer and the stiff sand layer below.
The larger the thickness of soil elements, the
higher the damage level of the piles especially
in the lower part of them. However, compared
with the 3-D analysis results shown in Fig 8, the
damage in the piles were too severe in the
results by Type B and C, and Type A mesh
could give a reasonable damage state.
Fig 14 shows the curvature distributions of
the left, center and right piles when the
maximum and minimum displacements were
reached at the top support. The difference by
the soil thickness in the 2-D models was
remarkable in the case with original input wave
rather than that with the moderate wave.
Comparing with the 3-D analysis results, Type
A mesh was found to give a reasonable result.
Consequently, from the viewpoint of the
deformation of pile foundation in these results,
Type A mesh in which the soil thickness was
equal to the footing width could give an
equivalent response to the 3-D model. In the
previous discussion on the soil deformation
around the pile foundation, the effective width of
soil to the response of foundation was found to
be about the width of footing as shown in Fig 10.
This tendency was consistent with the 2-D
analysis result in this section. Therefore, it was
suggested that the soil thickness in 2-D model
should be determined according to the effective
region of the soil-structure interaction in each
analytical target.

416

In this paper, 3-D and 2-D seismic response


analyses of RC pile-soil system were conducted
using nonlinear finite element method, and the
effect of the thickness of soil elements used in
the 2-D models on the response of pile
foundation was investigated by comparing
these analytical results. In the results obtained
by the 2-D models with different values of soil
element thickness, macroscopic response at
the top support was similar, whereas the
damage level in the RC piles was remarkably
different. For the target structure in this paper,
the 2-D model in which the thickness of soil
elements was equal to the footing width could
give an equivalent result to the 3-D model.
Finally, the importance was emphasized that
the thickness of soil elements in 2-D finite
element models should be carefully determined
according to the effective region of the soilstructure interaction in each analytical target.
REFERENCES
Japan Society of Civil Engineers (2005): Standard
Specifications for Concrete Structures-2002
"Seismic Performance Verification", JSCE
Guidelines for Concrete No.5, JSCE. (Original
specification written in Japanese was published
in 2002.)
Ishihara, T. and Miura, F. (1994) Comparison of
Seismic Responses of 3-Dimensional StructurePile Foundation-Ground Interaction Systems with
Those of 2-Dimensional Systems, Journal of
Structural
Mechanics
and
Earthquake
Engineering, Vol. 501, No.I-29, pp.123-131.
Kimura, M. and Zhang, F. (2000): "Seismic
Evaluations of Pile Foundations with Three
Different Methods based on Three-Dimensional
Elasto-Plastic Finite Element Analysis", Soils and
Foundations, Vol.40, No.5, pp.113-132.
Okamura, H. and Maekawa, K. (1991): Nonlinear
Analysis and Constitutive Models of Reinforced
Concrete, Gihodo Shuppan, Tokyo, Japan.
Maekawa, K., Pimanmas, A. and Okamura, H.
(2003): Nonlinear Mechanics of Reinforced
Concrete, Spon Press, London.
Ohsaki, Y. (1980): "Some Notes on Masing's law
and Non-Linear Response of Soil Deposits",
Journal of the Faculty of Engineering, The
University of Tokyo (B), Vol.XXXV, No.4, pp.513536.

Seismic Behavior of Single Pile in Cohesive Soil


R. Tuladhar1, H. Mutsuyoshi2 , T. Maki3
1,2,3

Graduate School of Science and Engineering, Saitama University, Japan

Abstract
For the development of seismic response analysis method for a whole structural system, it
is first imperative to clarify the performance of pile under seismic loading. In this study, fullscale lateral loading tests were carried out on single concrete piles embedded into the
ground. The test piles were hollow precast prestressed concrete piles with outer diameter
of 45cm and thickness of 7cm. The length of the test piles were 14m. The experimental
observations showed that the soil gap occurs between soil and pile in the face opposite to
the loading direction. It further showed that significant degradation in lateral load carrying
capacity of pile occurs when subjected to cyclic loading. The experimental data were used
as the basis for 3D finite element analysis. The study showed that with proper
consideration of interface element between soil and pile surface and the degradation of
stiffness in soil due to cyclic loading, 3D finite element analyses can simulate the seismic
behavior of the concrete pile well.
INTRODUCTION

Regarding the analysis of soil-pile


interaction during earthquakes, there are
simplified approaches to analyze laterally
loaded piles which model pile as elastic and the
surrounding soil as discrete springs and
dashpots (Matlock 1970, Reese 1975). These
discrete
models,
however,
ignore
the
nonlinearity of piles and cannot model the
damping and inertial effects of soil media.
With the advancement in computation
capability, 3D finite element method (FEM) has
become more appealing, as in this method soil
can be modeled as continuum media taking into
account damping and inertial effects of soil. In
addition, the nonlinearity of pile and soil can be
taken into consideration more rigorously. There
are number of studies carried out to investigate
the behavior of pile and soil by FEM analysis
(Trochanis et al., 1991; Wakai et al. 1999; Maki
and Mutsuyoshi 2004). These studies, however,
lack proper calibration with the field
experimental studies. Moreover, there is still
room for research to address various issues like
interface between pile and soil and degradation
of lateral load carrying capacity of pile in cyclic
loading. Hence, the main purpose of this
research is to study the lateral behavior of
concrete pile and soil during earthquakes using
full-scale lateral loading tests and to use 3D

Damage sustained in recent earthquakes,


have highlighted that the seismic behavior of a
structures is highly influenced by the response
of the foundation and the ground (Mylonakis et
al., 2000). Hence, the modern seismic design
codes recommend the seismic response
analysis of an overall structure system including
soil-structure interaction. For the development
of seismic response analysis method for a
whole structural system, however, it is first
imperative to clarify the behavior of soil and pile
during earthquakes. Moreover, observations of
damages in pile foundations in recent
earthquakes have stressed the necessity of
detailed study on performance of pile under
seismic loading.
There have been number of studies carried
out on small scaled concrete piles embedded
into model ground (Maki and Mutsuyoshi,
2000). However, considering the complexity of
the interaction between pile and the supporting
soil, analysis of experimental data from fullscaled tests is imperative to understand the
actual behavior of piles during earthquakes. In
addition, valid verification of present analytical
tools cannot be made without well documented
data from the full-scale testing of the
instrumented piles in the field.

417

Experimental setups

FEM analysis to address these issues.


Full-scale monotonic and reversed cyclic
lateral loading tests were performed on two
instrumented concrete piles embedded into the
ground. 3D FEM analysis was then carried out
to study the behavior of the experimental
specimens. In 3D FEM analysis, soil and
concrete pile were modeled as 20-node solid
elements. The effect of use of interface element
between pile and soil, and degradation of lateral
load carrying capacity of pile in reversed cyclic
loading were also investigated in the study.

Piling was carried out by inside drilling


method. An auger of 270mm diameter was
inserted into the hollow pile and drilling was
carried out inside the pile while pile was being
embedded into the ground. The test piles were
embedded up to 12.8m from the ground level
(GL). The head of the pile and the loading point
were 1.2m and 0.6m from GL, respectively.
The standard penetration test (SPT) was
carried out at the experimental site to
investigate the relevant soil parameters. The
NSPT values obtained from the test are shown
in Fig. 2 along with soil type. The depth of
water-table was 1.3m from the ground level.
Reaction frames were setup using six
reaction steel piles as shown in Fig. 3. The
reaction piles were driven up to the depth of
10m. Experimental setups were done in
accordance with the JSF standards (JSF 1983).
Test pile SP1 was subjected to monotonic
loading, whereas test pile SP2 was subjected to
reversed cyclic loading.

FULL SCALE LATERAL LOADING TEST

The experimental program consists of lateral


loading test on two full scaled concrete piles
embedded into the ground. One of the test piles
was subjected to monotonic loading, whereas
another test pile was subjected to reversed
cyclic loading.
Details of test pile

Both of the test piles were hollow precast


prestressed concrete piles with a diameter of
450mm and a thickness of 70mm (Figure 1). 12
prestressing steel bars of 7mm diameter were
used for longitudinal reinforcements and spiral
hoop reinforcements with a diameter of 3mm
and pitch of 100mm were used as lateral
reinforcements. Strain gages were attached to
the longitudinal reinforcements as shown in Fig.
1. Compressive strength of concrete (fc) was
79MPa and yielding stress (fy) of longitudinal
prestressing steel was 1325MPa. Effective
prestress on concrete piles was 5MPa.
Bending test was carried out on an 8m long
pile with the same cross-sectional properties as
the test piles to investigate the sectional
properties of the test pile. From the bending
test, yielding moment and ultimate moment of
the section were obtained as My = 124 kNm and
Mu = 160 kNm, respectively.

N-SPT
0

10

20

30

Soil
Type

40
0

10

10

clay
silt

sand

Depth : m

silt and sand


silt
20

20

silt and sand


clay
30

30

Fig 2: Soil profile at test site

GL
2x10@0.3m

2x10@0.6m

2x3@1.2m

0.8m
12 no. 7mm
dia PC bar

0.6m

0.6m

3m

6m

3.6m

0.31m
0.8m
0.45m

14m

Fig 1: Test piles details and position of strain gages

418

In monotonic loading, a gap of 52mm was


observed in soil in active side (extension side)
at the maximum pile head displacement of
75mm. In reversed cyclic loading, gaps of
60mm and 75mm were observed between soil
and pile surface at each side of the pile at
maximum pile head displacement of 110mm.
Curvature distributions along the pile shaft
were calculated from the measured strain data,
as shown in Fig. 5(a) and Fig. 5(b). For
specimen SP1, the maximum curvature was
attained at the depth of 0.9m from GL, and in
specimen SP2 the maximum curvature was
observed at the depth of 1.5m from GL.
Damage in the piles below the ground
surface was examined by digging trench around
the piles after completion of the loading test. In
specimen SP1, maximum damage was
observed at 0.9m from GL, whereas, maximum
damage was observed at the depth of 1.5m
from GL in specimen SP2. From the
experiment, it was observed that the depth of
plastic hinge was lowered in the specimen
subjected to reversed cyclic loading compared
with that under monotonic loading. The lowering
of location of plastic hinge for specimen SP2
subjected to reversed cyclic loading compared
to specimen SP1 subjected to monotonic
loading is due to the degradation of soil
stiffness due to reversed cyclic loading.

Fig 3: Experimental setup


EXPERIMENTAL OBSERVATIONS

150

Curvature : 1/cm
0
0

Depth : m

Load displacement relationships for the test


piles SP1 and SP2 are shown in Fig. 4. In case
of monotonic loading, SP1, yielding of the pile
occurred at 120kN and maximum lateral load
carrying capacity was 135kN.
Here, the
specimen is assumed to yield when the
longitudinal bars at tension side reach yielding
strain. The maximum displacement at failure
load was 75mm.
For the specimen SP2 subjected to reversed
cyclic loading, maximum lateral load carrying
capacity was 114.5kN at the maximum
displacement of 110mm. Compared with the
monotonic loading, the lateral load carrying
capacity of the pile under reversed cyclic
loading has degraded significantly by 15%. The
degradation in lateral load carrying capacity in
reversed cyclic loading is due to the
degradation in shear modulus of soil with cyclic
loading. In both of the specimens SP1 and SP2,
failure occurred with breaking of PC tendon.
The failure modes for both of the specimen
were flexural failure with breaking of one of the
prestressing bars in tension side.

Load : kN

0.9

1.8

1.8

2.7

2.7
10mm

4.5

50

Curvature : 1/cm
0.0004-0.0003
-0.00015 0 0.000150.0003
0

0.9

3.6

100

0.0002

20mm

3.6

30mm

25mm

-25mm

40mm

40mm

-40mm

55mm

-55mm

70mm

-70mm

50mm

4.5

60mm

5.4

Fig 5: Curvature distribution (a) SP1 and (b) SP2

-50
SP1

-100
-150
-150

5.4

SP2

-100

-50

50

100

3D FINITE ELEMENT ANALYSIS


150

The three-dimensional FEM analysis was


used to calculate numerically the behavior of
the experimental specimens. In this modeling,

Displacement : mm

Fig 4: Load displacement curve for SP1 and SP2

419

both the pile and soil are modeled as 20-node


isoparametric solid elements. Soil properties
used in the analysis are shown in Table 1.

Normal
stress
Opening

Table 1: Soil properties used in the analysis


Depth
from
GL
0 6m
6m
12.8m

Clay

Unit
weight
(kN/m3)
15.7

Shear
strength
(kPa)
33

Shear
modulus
(MPa)
20.4

Sand

18.6

140

154.3

Soil
type

Normal
strain
1
Closure

Fig 7: Opening closure model for interface

Soil and pile were modeled up to 12.8 m


depth, and 9.5m and 3.15m in length and width,
respectively as shown in Fig. 6. Only half of the
domain was considered taking into account the
symmetry in the geometry and load. The base
was fixed in all X, Y, and Z directions. Two
lateral faces of soil model, perpendicular to the
direction of loading were fixed in X direction and
the remaining two lateral faces were fixed in Y
direction. To simulate the gap formation
between soil and pile surface, a 16-node
interface element is used between soil and pile
surface (Fig. 7). In this opening-closure model,
there is no stress transfer between pile and soil
during opening or tension. During closure or
compression, however, high rigidity (K) is
assumed to avoid the overlapping of soil and
pile element in compression. In this case, no
shear stress between RC and soil element is
assumed.

Pile

Kn

Constitutive model for concrete and soil

Nonlinearity of concrete before cracking was


modeled by elasto-plastic fracture model
(Maekawa et al., 2003). A smeared crack model
based on average stress-average strain was
used to model concrete after cracking. For post
cracking behavior compression and tension
model proposed by Maekawa et al. (2003) were
used. For reinforcements, nonlinear path
dependent constitutive model of Kato (1979)
was used.
Soil elements were formulated in both
deviatoric
and
volumetric
components
separately. The volumetric component of soil
element was taken here as linear elastic. For
the deviatoric component, non-linear path
dependency of soil in shear was modeled by
Ohsaki model (1980). This model consists of a
hyperbolic skeleton curve and hysteretic curve
is formulated based on Masings law (Ohsaki
model, 1980).
Material parameters of soil, initial shear
modulus G0 and shear strength Su, were
calculated from N-SPT value measured at the
field, by using Eq (1) and (2):

Loading
point

12.8mm

G0 = 11.76 N

0.8

(N/mm2)

(1)
2

Su = G0 1100 ( for Sand ) (N/mm )


= G0 600

( for Clay ) (N/mm )

(2)

Analysis parameters

3.15m

The analysis parameters are shown in Table


2. For monotonic loading, interface element
between pile and soil was not considered in
case Mon1. Whereas in case Mon2, 16-node
interface element was considered between pile
and soil. In case of reversed cyclic loading,
shear stiffness degradation of soil due to cyclic
loading was not considered in case Rev1.

9.5mm

Fig 6: Finite element model for pile and soil

420

Whereas in case Rev2, stiffness degradation


factor was considered.

Curvature : 1/cm
0

0.0002 0.0004 0.0006

Table 2: Analysis parameters

0.9

Loading

Name

Monotonic

20-node
solid
element

Reversed
cyclic

Mon1
Mon2
Rev1
Rev2

1.8
Depth : m

Pile
Element
Type

Descriptions
Stiffness
Interface
reduction
element
factor
No
-Yes
-Yes
No
Yes
Yes

20mm
30mm

3.6

40mm
50mm

4.5

60mm
80mm

FINITE ELEMENT ANALYSIS RESULTS

5.4

Fig 9: Curvature distribution case Mon2

Monotonic loading

Case Mon1 without interface element


between pile and soil highly over estimates the
lateral loading capacity of the pile as seen in
Fig. 8. In case Mon2 where interface element
has been considered, the load displacement
curve agrees well with the experimental results.
Curvature distribution along the depth of pile
for case Mon2 is shown in Fig. 9. The depth of
maximum curvature and curvature distribution
along the depth of the pile agrees well with the
experimental observations.
This suggests that the 20-node solid
element modeling of pile with interface element
between pile and soil can accurately simulate
the lateral behavior of pile and soil. Moreover,
the formation of gap between soil and pile can
be simulated realistically by using the interface
element as in this study.

Reversed cyclic loading

Fig. 10 shows the load displacement curve


for reversed cyclic loading for case Rev1.
Interface element between pile and soil was
being considered in case Rev1, however, it still
overestimate lateral load capacity of the pile.
The degradation in shear stiffness of
cohesive soil due to cyclic loading has been
studied by means of undrained cyclic triaxial
and simple shear tests on clay by several
researchers (Thiers and Seed, 1968 and
Gerolymos and Gazetas, 2005). Shear modulus
was found to decrease approximately 50-80%
for peak strain greater than 1%.
The reduction in shear modulus of soil with
cyclic loading leads to reduction in the lateral
load carrying capacity of pile. In the current
Ohsaki model used in the analysis, the
deviatoric stress-strain relationship has been
tested only up to 1% engineering strain level.
After this strain, constant shear modulus (G2) is
assumed as shown in Fig. 11.

250
200
Load : kN

10mm

2.7

150

160

100

80

50

Load : kN

Experiment (SP1)
Mon1
Mon2

0
0

20

40

60

80

-80
Experiment (SP2)

Displacement : mm

Rev1

-160
-150

Fig 8: Load displacement curves from experiment


and analysis

-100

-50

50

100

150

Displacement : mm

Fig 10: Load displacement curves case Rev1

421

To incorporate the degradation in the soil


stiffness due to cyclic loading, the shear
modulus after 1% shear strain was reduced
parametrically using stiffness degradation factor
(Ks) considering the experimental result.

Curvature : 1/cm
-0.0006-0.0003 0
0

0.00030.0006

0.9

Depth : m

1.8
G2

Ks xG2

2.7
3.6

2G0
0

4.5

0.01

-12mm

12mm

-25mm

25mm

-40mm

40mm

-55mm

55mm

-65mm

65mm

5.4
Ks = Stiffness degradation factor

Fig 13: Curvature distribution case Rev2

Fig 11: Incorporating stiffness degradation factor in


Ohsaki Model

CONCLUSIONS

In case SL-Rev2 stiffness degradation factor


(Ks) of 0.2 was considered. For the stiffness
reduction factor of 0.2, the analytical results
tend to agree with the experimental results
(Fig.12). This shows that the degradation in
stiffness of soil due to reversed cyclic loading is
as high as 80%.
This implies that the
degradation in stiffness of soil due to reversed
cyclic loading should be considered to
accurately calculate the seismic behavior of
concrete piles.
Fig. 13 shows the curvature distribution
along the depth of the pile for case Rev2.
Maximum curvature depth is 1.2m from the GL.
The lowering of the depth of plastic hinge in
reversed cyclic loading compared to the
monotonic
loading
also
signifies
the
degradation of the stiffness in soil due to
reversed cyclic loading.

In this study, full-scale monotonic and


reversed cyclic lateral loading tests were
performed on two instrumented concrete piles
embedded into cohesive soil.
Threedimensional finite element analysis was then
carried out to study the behavior of the
experimental
specimens.
From
the
experimental study and 3D finite element
analysis, the following conclusions can be
drawn:
When pile embedded in cohesive soil is
subjected to lateral loading, gap occurs
between pile surface and soil.
The gap
formation significantly reduces the lateral
capacity of the pile. The interface element
should be considered between pile surface and
soil to incorporate the gaping effect. Interface
element used in this study can realistically take
into account the gap formation.
It was also revealed that significant
degradation in lateral load carrying capacity of
pile occurs under reversed cyclic loading
compared to the monotonic loading.
The
reduction observed was around 15% in the
experiment. The degradation in reversed cyclic
loading is due to the degradation of shear
stiffness of clay due to cyclic loading.
With proper consideration of constitutive
models of soil and pile, interface element
between pile surface and soil, and the
degradation of soil stiffness under cyclic
loading, the three-dimensional finite element
analysis was found to simulate well the actual

160

Load : kN

80

-80

Experiment (SP2)
Rev2

-160
-150

-100

-50

50

100

150

Displacement : mm

Fig 12: Load displacement curves case Rev2

422

Reese, LC. (1975) Lateral loading of deep


foundations in stiff clay, Journal of the
Geotechnical Engineering Division, Proceedings
of the American Society of Civil Engineers, Vol.
101, No. GT7, pp. 633-649
Trochanis, AM., Bielak, J. and Christiano, P. (1991)
Three-dimensional nonlinear study of piles,
Journal of Geotechnical Engineering, Vol. 117,
No. 3, pp. 429-447
Thiers, GR. and Seed, HB. (1968) Cyclic stressstrain characteristics of clay, Journal of the Soil
Mechanics and Foundation Division, Proceedings
of the American Society of Civil Engineers, Vol.
94, No. SM2, pp. 555-569
Wakai, A., Gose, S. and Ugai, K. (1999), 3-D
elasto-plastic finite element analysis of pile
foundations subjected to lateral loading, Soils
and Foundations, Japanese Geotechnical
Society, Vol. 39, No. 1, pp. 97-111

behavior of pile and soil. Degradation in lateral


capacity of pile in reversed cyclic loading can
be taken into account by considering the shear
stiffness reduction factor as used in this study.
ACKNOWLEDGEMENT

The FEM Code, Concrete Model in 3D


(COM3), was offered by Prof. Koichi Maekawa,
University of Tokyo for this research. The
authors sincerely acknowledge his kind
cooperation.
REFERENCES
Gerolymos, N. and Gazetas, G. (2005) Constitutive
model for 1-D cyclic soil behavior applied to
seismic analysis of layered deposits, Soils and
Foundations, Japanese Geotechnical Society,
Vol. 45, no. 3, pp. 147-159.
JSCE (2005), Standard Specifications for Concrete
Structures: Seismic Performance Verificatio,
Tokyo: Japan Society of Civil Engineers.
JSF (1983) Standard method for lateral loading test
for a pile - T 32-8, Japanese Society of
Foundation Engineering, Tokyo
Kato, B. (1979) Mechanical properties of steel
under load cycles idealizing seismic action,
Bulletin D Infomration CEB, AICAP-CEB
Symposium, Vol. 131, pp. 7-27, Rome
Maekawa, K., Pimanmas, A., and Okamura, H.
(2003), Nonlinear Mechanics of Reinforced
Concrete, Spon Press, New York
Maki T. and Mutsuyoshi H. (2000) Response
behavior of RC piles under severe earthquake,
Proc. of the 12th World Conf. on Earthquake
Engineering, Paper no. 1245, Auckland, New
Zealand
Maki, T. and Mutsuyoshi, H. (2004) Seismic
behavior of reinforced concrete piles under
ground, Journal of Advanced Concrete
Technology, Japan Concrete Institute, Vol. 2, No.
1, pp. 37-47
Matlock, H. (1970) Correlations for design of
laterally loaded piles in soft clay, Proceedings of
Offshore Technology Conference, Paper no.
OTC1204, pp. 577-594, Dallas, Texas
Mylonakis, G., Gazetas, G., Nikolaou, S. and
Michaelides, O. (2000) The role of soil on the
collapse of 18 piers of the Hanshin Expressway
in the Kobe Earthquake, Proceedings of the 12th
World Conference on Earthquake Engineering
[CD-ROM], Paper no. 1074, Auckland, New
Zealand
Ohsaki, Y. (1980) Some notes on Masings law and
non-linear response of soil deposits, Journal of
the Faculty of Engineering, The University of
Tokyo, Vol. 35, No. 4, pp. 513-536.

423

Passive and active stress wave techniques for damage assessment of


railway piers
T. Shiotani1, S. Miwa1, D. G. Aggelis1, X. Luo2, H. Haya2
1

Research Institute of Technology, Tobishima Corporation, Japan


2
Railway Technical Research Institute, Japan

Abstract
In the present paper the internal condition of a concrete railway pier is evaluated using
acoustic emission and seismic tomography approaches before and after repair. Analysis
based on the amplitude distribution of AE events and more specifically the improved-b
value, was successful to quantify the repair effect, while also the propagation velocity was
increased. Results are verified by visual observations of cores retrieved, borehole camera
images as well as fluorescent epoxy impregnation. It seems that Non Destructive
techniques supply global information that cannot be revealed by any other applicable
technique.

INTRODUCTION

The number of civil structures older than 50


years continuously increases. The maintenance
and renewal cost is expected to exceed that of
new construction in the following years.
Therefore, the determination of the structural
integrity of concrete has become a priority.
Crack density, as well as crack width, have
been treated as damage indices so far;
however, cracks are conventionally evaluated
on the basis of surface observation only.
Acoustic emission (AE) tests supply information
about the severity and location of damage
through the source location and waveform
parameter analysis. In this paper, the AE
testing is applied for both the damaged and
repaired state of a railway concrete pier. Nondestructive testing techniques can be used for
global health monitoring to provide information
for the location of damaged areas. Destructive
techniques, which include core retrieving and
visual observation, can provide a more accurate
insight but for a small area. Therefore, AE and
seismic tomography are performed in order to
monitor the pier in a global way, whereas the
results are followed by visual observation at the
points where cores were extracted.

MONITORING PROCEDURE
Piers of a concrete railway bridge, located in Chiba
prefecture were inspected. Herein results of a
representative case will be discussed. Details can
be found in (Shiotani et al. 2005b). The piers were
made from plain concrete, while the pier of interest
is pier 3, see Fig. 1(a). The structure has sustained
deterioration due to long-term weathering as well
as earthquakes. Years ago, the degree of damage
of each of the piers was estimated by impact
vibration tests (Kikuchi et al. 1988, Nishimura and
Haya, 1991) implying that the damage was more
severe for P3 compared to the other piers. In Fig.
1(b) the pattern of observed surface cracks on pier
3 is depicted. The horizontal pattern of cracks
corresponds to the construction joint of concrete,
while a part of concrete was already separated
from the stem of the pier.
AE tests were carried out using the excitation of
passing trains and were followed by seismic
tomography. Three cores were excavated in the
horizontal direction and one diagonal, as seen in
Fig. 1(c). A borehole CCD camera was inserted to
examine the crack distribution and the cores
retrieved were impregnated with fluorescent epoxy.
This way crack widths smaller than 1m can be

424

revealed under the ultraviolet light (Shiotani et


al. 2002).

(a)

(b)

by the load of passing train, were detected with the


sensors, amplified and stored in a Mistras,
(Physical Acoustics Corp.) monitoring system.
Accelerometers (ASW-2A, max acceleration 2g,
Kyowa Dengyo Corp), -shaped displacement
meters (PI-2-50, Kyowa Electronic Instruments)
and strain meters (PL-120, Kyowa Dengyo Corp)
were also applied, with sampling rate of 1 kHz and
stored in a data logger (DRA-101B, Tokyo Sokki
Kenkyujo).

(c)

Unit: mm

Fig 2: Arrangement of AE sensors with other


deformation-sensors.
Fig 1: (a) Railway concrete piers tested, (b) Crack
condition of pier 3. (c) locations of excavated
boreholes.

In the specific case of pier 3, repair was


conducted by impregnation of inorganic agent
CS-21 with fine cement particles. Before
injection, all visible cracks were sealed with
caulking material. The injection pressure,
applied by a rubber strip, varied according to
the amount of injection agent. After injection,
AE and seismic tomography tests, were
repeated in order to evaluate the repair effect.

Seismic Prospecting
Seismic tomography was performed for three
sections as shown in Fig. 3. In each section, 12
piezoelectric
accelerometers
(SAF51,
Fuji
Ceramics Corp.), symmetrically arranged were
placed onto the pier surface spaced by 0.5m. In the
seismic prospecting, elastic waves were excited by
the impact of a 50mm diameter steel ball. The
velocity structure in the section was calculated by
means of ray-tracing and simultaneous iteration
methods (Kobayashi et al. 2006).

DESTRUCTIVE MEASUREMENTS

EXPERIMENTAL DETAILS
Twelve AE sensors of 60kHz resonance
were placed on the pier surface, as seen in Fig.
2. They were placed at four heights in a set of
three, while the vertical spacing was 1m. AE
monitoring was performed when trains passed
over the bridge. AE events that were induced

Using borehole image-processing system, BIPS


(Yasuda et al. 1992), optical images of the interior
of the boreholes, were recorded. The BIPS enabled
to obtain quantitative/qualitative results by
projected/unfolded image data. Besides the
borehole image monitoring, cores retrieved were
carefully observed visually to quantify the damage.

425

smaller than 0.5mm can be observed. The depth of


1.4m corresponds to that of the construction joint of
concrete. Beside the observed crack and a void at
0.4 m depth, no other cracks were found. The core
samples excavated at that points are also shown in
Fig.4 (right side). A large number of ruptures,
especially below 1.6m depth, were obtained. Since
cracks corresponding to those depths were not
observed in the borehole image, it is likely that
these ruptures were induced to the cores during the
excavation.

Fig 3: Arrangement of accelerometers for seismic


prospecting in P3. (top: side view, bottom:
plan view)

Resolution limitations or possible damaging


of the cores during excavation presents a
difficulty in distinguishing real cracks from
apparent cracks. Therefore, an impregnation
technique with fluorescent epoxy is used to
enhance the characterization. In this test, the
bulk cores retrieved were placed into an
impregnation chamber filled with fluorescent
epoxy, under a reduced pressure (9.3 Pa) using
a vacuum pump. By sectioning, cracks
distributed within the core can be visualized
under the ultraviolet light (Shiotani et al. 2002).
Note that sectioning was conducted after the
impregnation and no additional cracks were
introduced.

RESULTS

Fig 4: Developed borehole image (left) and core


retrieved (right).

Additionally, the core samples ranging from 1.11.6m were subjected to epoxy impregnation, and
cuts were made at four locations represented as 14 as shown in Fig. 4. Results are given in Fig. 5;
there one side of the cross-section at No. 1 and No.
4 are shown. Under ultraviolet light, No. 1 section
includes voids of large area. At No. 4, which is
located close to the casting joint, a distinct macrocrack and branched cracks were detected. These
findings imply that even for the bulk core/unruptured core, large scale of cracks and large area
of voids existed, so that it is most likely that these
critical cracks can readily generate AE activity due
to a train induced load.

Observation in Damaged Condition

An example of a developed borehole image


is depicted in Fig. 4. At 1.4m depth, a crack

426

(c) repaired state, side view (d) repaired state, top


view.

(a)

Fig 5: Sections after the impregnation at No. 1 (top


left) and No. 4 (bottom left), and the
corresponding images under ultraviolet light.
(No. 1: top right, No. 4: bottom right).

Acoustic Emission
As to pier 3, AE sources in three dimensions
are shown in Fig.6(a) and (b). The diameter of
AE source reflects the average of peak
amplitude of AE hits contributing to the AE
source/event. No significant AE sources were
obtained in pier 2; however, substantial AE
sources were detected both in P3 and P4 were
found. These extended

(a)

(b)

(c)

(d)

not only along the existing joints but in the whole


monitoring area. The 3D AE sources after repair
are shown in Fig. 6(c) and (d). Comparing to the
damaged condition, only slight decrease of the
scale of AE sources as well as the number of AE
sources was found in the repaired piers. The
average of the peak amplitude of the events was
39.82dB, before repair and 39.24dB after repair. It
can be seen that after repair the amplitudes were
more uniformly distributed.
In general however, notable change or
improvement due to repair could not be obtained by
the simple examination of AE events amplitude.
Concerning the large cracks of the construction
joint, these could be observed by the CCD camera.
However, as seen, the AE sources expand to wider
areas of the structure. This means that distributed
critical cracks were present and with the additional
torsion applied by the boring, they grew to
macroscopic cracks rupturing the cores, see Fig. 4.
This was verified by the results of epoxy
impregnation. Therefore, distributed critical cracks
were present in the structure which could not be
identified by conventional techniques like borehole
camera and unaided eye observation. This was
possible with AE however, by locating events,
which only epoxy impregnation could reveal.
Quantification of Damage using the improved bvalue analysis
In order to evaluate structural integrity, Calm ratio
and RTRI have been proposed on the basis of
train-induced AE activity, and they were
successfully applied for damage evaluation of
railway substructures (Luo et al. 2004). When
obtaining these indices, structural behavior such as
deformation is essential, since the loading as well
as the unloading process should be accurately
defined. This can be determined in a laboratory
experiment with a simple loading cycle; in the
present case however, a specific point could not be
defined between loading and unloading due to
complexity of strain curves, see Fig. 7.

Fig.6: AE sources in three-dimension, (a) damaged


state, side view, (b) damaged state, top view,

427

Fig 7: Symmetric behavior of two -shaped


displacement meters along with AE activity
due to train passage.

In that respect the AE amplitude distribution


was exploited to yield a unique trend
corresponding to the degree of damage.
Cumulative amplitude distributions are shown in
Fig. 8. This was done in the basis of the
improved b-value analysis (Shiotani et al. 1999,
Shiotani et al. 2001, Shiotani et al. 2004,
Shiotani et al. 2005a).
After an AE event is generated, and while
propagating into the material, it is certainly
influenced by the material condition. Especially
in a damaged state, it is certain that any waves
undergo multiple reflections and scattering on
cracks, discontinuities or voids, other than their
own source. Therefore, although in a severely
damaged situation a critical crack is expected to
emit events with high intensity, it is not certain
that the AE sensors will detect high amplitude
events due to the resulting attenuation.
However, it has been found that in severely
damaged situations the proportion of higher
amplitudes within a distribution is increased
compared to the lower amplitudes. In intact
state, the amplitude of events is more uniformly
distributed (Shiotani et al. 1999, Shiotani et al.
2001). In this case, the distribution of event
amplitudes and not solely the average
amplitude can lead to useful quantification. As a
typical result, the cumulative amplitude
distribution in P3 is shown in Fig. 8. In the state
of before repair, or damaged condition, more
AE events of large amplitude are generated,

while in the after repair state, more small amplitude


events are generated. Such a difference of
distribution can be quantified using the improved bvalue. Therefore, the gradient of the curve is
extracted. This gradient is 0.0817 in the damaged
condition and 0.2508 in the repaired condition for
pier 3. This difference is distinct enough to separate
the two conditions. According to empirical data, Ib
values higher than 0.2 imply minor damage or
intact state, while values lower than 0.1 imply
serious damage. The improved b-value analysis
has already been applied to several railway
structures and was successful in quantifying the
degree of damage. Note that when comparing a
seismic b-value, the improved b-value should be
multiplied by a coefficient of 20.

Fig 8. Amplitude distributions in P3. The average


amplitude of a set of AE hits contributing to an AE
event is used to determine the distribution.

Seismic Prospecting
A typical distribution of elastic wave velocity on a
section before repair is shown in Fig. 9(a). together
with 3D AE sources projected to this section. It is
evident that low velocity areas of less than 3000m/s
are exhibited near the construction joint. However,
low velocity is observed even in areas where there
is no influence by the large crack at the joint where
significant AE events were detected. In Fig. 9(b) the
corresponding tomogram after repair is depicted. It
seems that the defect near the joint was sufficiently
eliminated and wave propagation was facilitated.
The agent was injected by the diagonal borehole 4
that was opened at the height of 2m. Since no
pressure was applied it is reasonable that areas
near or above 2m do not exhibit similar increase.

428

and the improved b-value demonstrate the


possibility to quantify damage and repair effect.
REFERENCES

Fig 9: Elastic wave velocities estimated by seismic


prospecting in P3. (left: before repair, right:
repaired, the color scale is in terms of m/s).

The effectiveness of repair work was thus


verified with seismic prospecting, while, as
analyzed earlier with the I-b value, acoustic
emission amplitudes were more uniformly
distributed in many moderate sources than
some specific high intensity sources. This is
also characteristic of the improvement.

CONCLUSIONS

Damaged railway concrete piers were


investigated with the AE technique before and
after repair. Details of damage condition were
also
studied
with
several
observation
techniques. Through those studies the following
conclusions are drawn:
1. AE activity reflects the actual damage
condition of structures. Using the improved bvalue, a quantification measure between
damaged and repaired state can be obtained.
2. Quantification of damage only from surface
visual observations and/or borehole wall
observation is not sufficient. Critical damage
assessment could only be done by means of
fluorescent epoxy impregnation, but this is
impractical in field.
3. Non-destructive techniques as AE and
seismic prospecting showed reasonable results
corresponding to the damage obtained from all
observations. Seismic prospecting can highlight
the effectiveness of repair works as an increase
of the wave velocity. It is expected that
attenuation
tomography
would
enhance
characterization since energy parameters are
more sensitive to existing inhomogeneity (Shah
et al. 2000, Streeter et al. 2003). AE analysis

Kikuchi, Y., Nishimura, A., and Yamada, M., (1988)


Investigation on the method of judging of integrity
for the foundation of the bridge, Proceedings of the
43rd JSCE Annual Meeting, Japan Society of Civil
Engineering, pp. 250-251, in Japanese
Kobayashi, Y., Shiojiri, H., and Shiotani, T., (2006)
Damage identification using seismic travel time
tomography on the basis of evolutional wave velocity
distribution model, Structural Faults and Repair2006, June 13-15, Edinburgh, (in CD-ROM).
Nishimura A. and Haya, H., (1991) Assessment of the
structural integrity of bridge foundation by the impact
vibration test, Proceedings of GEO-COAST 91, pp.
719-724
Luo, X., Haya, H., Inaba, T., Shiotani T., and Nakanishi,
Y., (2004) Damage evaluation of railway structures
by using train-induced AE, Construction and
Building Material, Vol. 18, 215-223
Shah, S. P., Popovics, J. S., Subramanian, K. V., and
Aldea, C. M., (2000) New directions in concrete
health monitoring technology, J. Eng. Mech.-ASCE,
V. 126, No. 7, pp. 754-760
Shiotani, T., Fujii, K., Aoki T., and Amou, K., (1999),
Evaluation of progressive failure using AE sources
and improved b-value on slope model tests,
Progress in AE VII, JSNDI, pp. 529-534
Shiotani T., Li, Z., Yuyama S., and Ohtsu, M., (2001),
Application of the AE
improved b-value to
quantitative evaluation of fracture process in
concrete-materials, Journal of Acoustic Emission,
Vol. 19, pp. 118-133
Shiotani, T., Bisschop, J., and van Mier, J.G.M. (2002)
Drying Shrinkage Microcrack Measurements in
Cementitious Composites Using AE and FLM,
Society of Experimental Mechanics, Proceedings of
the 2002 SEM Annual Conference & Exposition on
Experimental and Applied Mechanics, Paper
Number: 202 (CD-ROM)
Shiotani, T., Nakanishi, Y., Luo X., and Haya, H., (2004),
Damage assessment in railway sub-structures
deteriorated using AE technique, 26th European
Conference on Acoustic Emission Testing (EWGAE
2004), Vol. 1, p. 255.

429

Shiotani, T., Luo, X., Haya H., and Ohtsu, M.,


(2005a) Damage quantification for concrete
structures by improved b-value analysis of AE,
11th International Conference on Fracture
(ICF11), (in CDROM)
Shiotani T., Nakanishi, Y., Iwaki, K., Luo, X., and
Haya, H., (2005b) Evaluation of reinforcement
in damaged railway concrete piers by means of
acoustic emission, Journal of Acoustic Emission,
Vol. 23, pp. 260-271
Streeter, K., Schuller, M., and Xi, Y., (2003)
Ultrasonic attenuation tomography of concrete
Engineering
Mechanics
structures,
16th
Conference (ASCE) Seattle, USA, July 16-18
Yasuda, T., Taniguchi, S., Kamewada S., and
Okano, G., (1992) Crack evaluation system for
analyzing the condition of superannuated tunnel,
Proceedings of the international symposium on
Fracture and Jointed Rock Masses, pp. 249-255

430

Liquefaction Densification Related Studies,


and others

Centrifuge Tests on Remedial Measure Using Batter Piles Against


Liquefaction-Induced Soil Flow After Quay Wall Failure
T. Tazoh1, M. Sato2, J. Jang1, and G. Gazetas3
1

Institute of Technology, Shimizu Corp., Japan, 2 National Research Institute for Earth Science & Disaster
Prevention, Japan, 3 National Technical University of Athens, Greece

Abstract
We experimentally investigate the behavior of a structure-footing-four piles system under
the action of seismic shaking causing liquefaction of a critical soil layer and subsequent soil
flow towards a laterally displacing quay-wall. This is a further study on our previous paper
st
presented at the 1 GreeceJapan Workshop on Seismic Design, Observation and Retrofit
of Foundations in 2005. In this paper, we proposed batter piles as a remedial measure, and
examined its effectiveness by means of centrifuge tests. A series of innovative centrifuge
tests explore in parametric fashion the role of the remedial measure in the response of the
system. The response is recorded in 38 channels providing time histories of acceleration,
displacement, excess pore water pressure, pile axial and bending strains, and earth
pressure. Five tests are carried out by changing the distance from the quay-wall to the pilefoundation-structures and the pile inclination angle. The results provide significant
quantitative information and qualitative insight on the behavior of this realistically complex
soil-foundation-structure system.

INTRODUCTION

mitigate the damage, is a crucial earthquake


engineering issue. (Tazoh, et al., 1996)
A series of centrifuge tests was conducted
to shed light on the seismic performance of pilefoundation-structure systems located behind
quay-walls, arising from the large displacement
or collapse of these walls and the ensuing soil
flow (Tazoh, et al., 2005).
Batter piles can be used with little or no
additional expense, no special design, and
hardly any difficulty in construction. In this
paper, remediation using batter piles is
proposed, and the effectiveness of this measure
is examined parametrically.

Soil liquefaction may cause significant


damage to structures, especially when lateral
soil flow takes place. Such flow (which often
takes the form of lateral spreading) was
triggered along river banks and sea coasts
behind quay-walls in Kobe during the 1995
Great Hanshin Earthquake. (Tazoh, et al., 2001,
2002)
Numerous studies have been published in
the last decade on lateral flow failures induced
by soil liquefaction. Numerical and experimental
studies have focused on the damage
mechanism of structures and on the external
forces induced by the soil flow on deep
foundations. Many important structures exist
along and near waterfronts in Japan, and their
safety in the next major earthquake must be
secured. Thus, the study of liquefaction-induced
soil flow, together with remedial measures to

CENTRIFUGE TESTING

To
investigate
quantitatively
the
effectiveness of remedial measures, we
compare the behavior of two structural systems:
with and without remediation respectively. Each

431

shaking table test for each model is carried out


under nearly identical conditions with respect to
input motions, soil liquefaction, and quay-wall
collapse. Note, however, that it is impossible to
achieve complete similarity between tests, due
to the difficulty of reproducing failure events of
soil and quay-walls.
Therefore, some innovative ideas must be
adopted in the centrifuge testing. The
experimental model used in the tests is shown
in Figure 1. In this study, a partition is placed at
the center of a soil container with two models in
each side: a vertical-pile-foundation-structure
and a batter-pile-foundation-structure. A quaywall consists of sheet piles. The pile-foundationstructures are installed parallel to the partition
behind a quay-wall with free tip-end
(unattached to the base), which will be
collapsed in the tests. The difference between
the responses of the two pile-foundationstructures represents the effect of the
remediation on the seismic performance of the
system. In these tests the loading is of a
kinematic nature: quay-wall collapse induces
soil flow failure, thereby loading the pilefoundation-structure system.

The thickness and relative density of each soil


layer are shown in Table 2. Cases 33-1 and 332 are tests carried out under the same
conditions: namely, the same distance from the
quay-wall to the pile foundations and the same
pile inclination angle, in order to investigate the
variability of these test results.

A laminar box is used as a soil container to


allow shear deformation of the deposit as in the
free field. The soil container is divided into two
parts: A-side and B-side where the vertical
pile foundation and the batter pile foundation
are installed, respectively. Photograph 1 shows
the model of the batter pile foundation used in
the tests.
Thirty-eight monitoring channels were
installed, with the sensors comprising
accelerometers,
pore
water
pressure
transducers,
strain
gauges,
non-contact
displacement meters, and earth pressure cells.
Moreover, numerous colored beads were
embedded in the soil layers to visually identify
the soil deformation at the final stage. All tests
were conducted at centrifugal acceleration of 30
g on a 1/30-scale model.

Fig 1: Longitudinal sections and plan of the 1/30scale centrifuge model of Case 32 (scale
unit: mm, for the prototype dimensions:
multiply by 30)

Five tests are carried out by changing the


distance from the quay-wall to the structure.
Table 1 shows the distances of these cases.
The soil deposit consists of four soil layers, of
which the first, third, and fourth soil layers are
non-liquefiable and the second is liquefiable.

432

Table 1: Five tests of the remedial measure with


batter piles
Case

Distance from quay-wall to

Inclination angle of

pile-foundation

batter piles

Table 3: Maximum absolute values of the input


motions and acceleration records in soil
layers
Case 31

200mm (prototype: 6m)

pile foundation)

pile foundation)

GL-20mmAG4

198

222

GL-90mmAG2

207

220

195

Case 32

100mm (prototype: 3m)

10

Case 33-1

50mm (prototype: 1.5m)

10

GL-130mmAG1

Case 33-2

50mm (prototype: 1.5m)

10

Input MotionAG0

Case 34

100mm (prototype: 3m)

Table 2: Soil profiles of the soil layers


Thickness

Soil profile

1st soil

70mm

Silica sand

Non- liquefiable

layer

(2.1m)

No. 8, Dr=50%

(above water table)

2nd soil

120mm

Silica sand

Liquefiable

layer

(3.6m)

No. 8, Dr=50%

3rd soil

80mm

Toyoura sand

layer

(2.4m)

GL-20mmAG4

Liquefiable

174
200

Case 32

Soil layer

B-side (Batter

10

(degree)
Case 31

A-side (Vertical

A-side (Vertical

B-side (Batter

pile foundation)

pile foundation)

215

209

GL-90mmAG2

508

212

GL-130mmAG1

220

244

Input MotionAG0

297

Case 33-1

A-side (Vertical

B-side (Batter

pile foundation)

pile foundation)

Dr=90%

GL-20mmAG4

173

218

GL-90mmAG2

216

172

GL-130mmAG1

201

4th soil

30mm

Silica sand

layer

(0.9m)

No. 3

Non- liquefiable
Non- liquefiable

Input MotionAG0

( ): prototype dimensions

236
205

Case 33-2
GL-20mmAG4

A-side (Vertical

B-side (Batter

pile foundation)

pile foundation)

175

227

GL-90mmAG2

148

195

GL-130mmAG1

177

152

Input MotionAG0

206

Case 34

A-side (Vertical

B-side (Batter

pile foundation)

pile foundation)

GL-20mmAG4

172

170

GL-90mmAG2

140

145

GL-130mmAG1

210

Input MotionAG0

255
203

Photograph 1: Batter pile foundation model used


in the tests

unit: gal

EFFECT OF THE REMEDIAL MEASURE AGAINST


LIQUEFACTION-INDUCED SOIL FLOW

Table 3 shows the maximum absolute


values of the input motions and acceleration
records in the soil layers. Figure 2 illustrates the
input motion (acceleration) of Case 32. Note in
Figure 2 that the time axis from 15 to 1200
seconds is at 1/8 of the scale from 0 to 15
seconds.

Acceleration(gal)

600
400

AG0(x)

200
0
-200
-400
-600
0

10

15
time(s)

400
20

800
25

Fig 2: Input motion (acceleration) of Case 32

433

1200
30

500
400

Quay-wall

300
200

Vertical-pile foundation

100

Batter-pile foundation
0

10

400
20

15
time(s)

800
25

Displacement(mm)
(Disp-AF,-BF)

Displacement(mm) (Disp-Y)

1200
1000
800
600
400
200
0
-200

-100
1200
30

Fig 3: Time histories of the horizontal displacements of the quay-wall (Disp-Y), and the footings of the vertical
piles in the A-side (Disp-AF) and of batter piles in the B-side (Disp-BF) of Case 32

given the significant difference between the


factors eff of the same condition tests of Cases
33-1 and 33-2, we should note that these
results may need to be interpreted with a
degree of caution.

Figure 3 shows the time histories of the


horizontal displacements of the quay-wall (DispY), and the footings in A-side (Disp-AF) and Bside (Disp-BF) of Case 32. Note that the axes in
the left and right sides in Figure 3 show the
displacements of the quay-wall and footings,
respectively. The effectiveness of the batter
piles in reducing the horizontal displacements of
the foundation is clear.
Photograph 2 shows the condition after the
test of Case 32. The displacements (residual
displacements) of the quay-wall and the
footings of the vertical pile foundation and the
batter pile foundation at 1200 seconds are
1048mm, 161mm, and 17mm, respectively. The
effectiveness of the batter pile foundation in
reducing the residual horizontal displacement of
the footing can be defined as the following
equation:

1
10
5

eff

0.8
0.6

Case33-2
Case31

0.4
Case33-1

0.2

Case34
Case32

0
0

2
4
6
Distance from Quay-wall (m)

Fig 4: Effectiveness of batter piles in reducing the


residual displacement of the footing

( Disp AF) - (Disp BF)


( Disp AF)
161mm 17mm

= 1
= 0.11
(at 1200 sec)
161mm

eff = 1

It means that the residual horizontal


displacement of the vertical pile foundation can
be reduced to 11% via the use of batter piles.
The factor eff of five tests can be summarized
as Figure 4. In this Figure, the vertical and
lateral axes show the factor eff and the
distance from quay-wall to the footing,
respectively. It is clear from the figure that
batter pile-foundations are effective in
restricting the horizontal displacement of the
foundation in a liquefied soil flow environment.
The factoreff of Case 32 with piles of 10
degree inclinations is smaller than that of Case
34 with piles of 5 degree inclinations. However,

B-side
A-side

Photograph 2: Post-test condition (Case 32)

434

0
A-S6, B-S6

-600
0

10

15
Time(sec)

20
400

25
800

Axial Strain()

300
0
-300

A-S7, B-S7

-600

Axial Strain()

10

15
Time(sec)

20
400

25
800

300
0
-300

A-S9, B-S9

-600
0

10

15
20
400
Time(sec)

25
800

0
A-S10, B-S10

-600
5

10

15
Time(sec)

20
400

25
800

10

15
Time(sec)

20
400

25
800

30
1200

Rear

600
300

Front

A-S2, B-S2

0
-300
-600
0

10

15
Time(sec)

20
400

25
800

30
1200

600
A-S4, B-S4

300

A-S6
A-S7

A-S1

A-S8

A-S3

A-S9

A-S4

A-S10

A-S5

A-S2

0
-300

Rear

Front

-600
0

Axial Strain()

300

-600

30
1200

600

-300

0
-300

30
1200

600

Vertical Pile
Batter Pile

A-S1, B-S1

30
1200

600

Axial Strain()

300

Axial Strain()

-300

600

Axial Strain()

300

Axial Strain()

Axial Strain()

600

10

15
Time(sec)

20
400

25
800

30
1200

600
A-S5, B-S5

300
0

B- EP

B-S1

B-S6
B-S7

B-S2

B-S8

B-S3
B-S4

B-S9

-300

10

-600

10

B-S5

B-S10

30
1200

10

15
Time(sec)

20
400

25
800

30
1200

1000
0
A-S6, B-S6

-2000

Bending Strain()

20
400

25
800

0
A-S7, B-S7

-1000
-2000
5

10

15
Time(sec)

20
400

25
800

1000
0
A-S8, B-S8

-2000
5

10

15
Time(sec)

20
400

25
800

1000
0
-1000

A-S9 B-S9

-2000
0

10

15
Time(sec)

400
20

800
25

A-S10, B-S10

0
-1000
-2000
0

10

15
20
Time(sec) 400

25
800

30
1200

A-S1, B-S1

-2000
0

10

15
20
Time(sec) 400

25
800

30
1200

2000

Rear

Front

1000
0
-1000

A-S2, B-S2

-2000
0

10

400
15
20
Time(sec)

800
25

1200
30

2000
1000

A-S6
A-S7

A-S1

A-S8

A-S3

A-S9

A-S4

A-S10

A-S5

A-S2

0
-1000

A-S3, B-S3

Rear

-2000
0

10

15
20
400
Time(sec)

25
800

Front

30
1200
B- EP

B-S1

B-S6
B-S7

2000
1000

B-S2

B-S8

0
-1000

A-S4, B-S4

B-S3
B-S4

B-S9

-2000

1200
30

2000
1000

0
-1000

30
1200

2000

Vertical Pile
Batter Pile

1000

30
1200

2000

-1000

2000

30
1200
Bending Strain()

15
Time(sec)

1000

0
Bending Strain()

10

2000

Bending Strain()

Bending Strain()

Bending Strain()

Bending Strain()

-1000

Bending Strain()

2000

10

0
Bending Strain()

Bending Strain()

Fig 5-1: Time histories of the axial strains of the piles obtained from the test (Case 32)

10

15
20
400
Time(sec)

25
800

30
1200

10

B-S10

2000
A-S5, B-S5

1000
0
-1000
-2000
0

10

15
Time(sec)

20
400

25
800

30
1200

Fig 5-2: Time histories of the bending strains of the piles obtained from the test (Case 32)

435

B-S5

Figure 5 compares the time histories of the


axial and bending strains in the vertical and
batter piles for Case 32, where the absolute
maximum values of the axial strains of the
batter pile foundation exceed those of the
vertical pile foundation. This must be caused by
the higher pushdown and pull-out forces
induced by the existence of the batter piles. On
the other hand, the bending strains at the pile
heads of the batter piles became smaller than
those of the vertical piles, clearly after the
excitation. This must be due to the footing with
batter piles rotating in the opposite direction to
that with vertical piles when the footing moves
laterally, as shown in Figure 6 (Tazoh et al.,
1987). This rotation of the footing allows the
bending strains at the pile head of the batter
piles to be reduced.
Figure 7 shows the factors obtained from
the division of the axial and bending strains of
the batter piles by those of the vertical piles,
respectively. In Figure 7(a) and (b), the factors
for the maximum and residual values of the
axial strains at the pile heads respectively are
mostly higher than 1.0, meaning the values of
the batter piles exceed those of the vertical
piles. Based on Figure 7 (a)~(c), the values of
the batter piles are almost higher than those of
the vertical values except for the residual
bending strain at the pile head in Figure 7 (d).
The reason why the factors become less than
1.0 is precisely because the footing with batter
piles rotates when the footing moves laterally,
as shown in Fgure 6.

(a)

Factor of axial strain

3
Maximun Value
2.5
2
Front

1.5

Rear

1
0.5
0
0

Distance from Quay-wall

(b)

Factor of axial strain

Residual Value

2.5
2
Front

1.5

Rear

1
0.5
0
0

Distance from Quay-wall

(c)

Factor of bending strain

2
Maximun Value
1.5
Front

Rear

0.5
0
0

Distance from Quay-wall

(d)

Factor of bending strain

2
Residual Value
1.5
Front

Rear

0.5
0
0

Distance from Quay-wall

Fig 6: Direction of the rotation of the footing with


batter piles when the footing moves laterally (Tazoh
et al., 1987)

Fig 7: Factors obtained from the division of the axial


and bending strains of the batter pile by those of the
vertical pile at the pile head

436

EARTH PRESSURE CAUSED BY


LIQUEFACTION-INDUCED SOIL FLOW

horizontal displacement (Disp-EF, -FF) of the


footings (pile caps) and the earth pressure (EP)
acting on the back of the footings, for Case 32.
Apparently, earth pressures acting on the
back of the footings (pile caps) increase during
excitation. This increase is huge in the case of
batter pilesan expected outcome in view of the
rigidity of this pile system and the consequent
diminution of lateral displacements. With
exclusively
vertical
piles,
the
lateral
displacements are nearly 10 times larger and
hence the earth pressures are substantially
smaller. Immediately after the end of excitation
the earth pressures reduce and finally they
completely disappear after 15sec, for both
batter and vertical piles.

Our measurements allow the study of earth


pressures generated by laterally flowing
liquefied soil, on the back of the foundation.
Many researchers have attempted to clarify the
mechanics of earth pressure after liquefaction
following the 1995 Kobe Earthquake. The
Japanese Highway Bridge code was revised in
1996 based on substantially increased earth
pressures.
On the contrary, other researchers have
claimed that the damage was caused by the
loss of the horizontal bearing capacity of the
soil, after failure of quay-wall, rather than from
the increased earth pressures (Tazoh et al.,
2002).
Figure 8 compares the time histories of the

Vertical-pile foundation

40
30
20
10
0
-10

60
50
40
30
20
10
0
-10

10

Batter-pile foundation

10

70
60
50
40
30
Horizontal displacement of footing
20
10
Earth pressure
0
-10
800
1200
15
20
25
30
400
time(s)

Displacement(mm)
(Disp-BF)

Earth Pressure (kPa)

300
250
Horizontal displacement of footing
200
150
100
50
Earth pressure
0
-50
800
1200
15
20
25
30
400
time(s)

Displacement(mm)
(Disp-AF)

Earth Pressure (kPa)

50

Fig 8: Comparison between the horizontal displacement of the footings and earth pressures
acting on the side of the footings (Case 32)

437

CONCLUSIONS

The main conclusions of the study are:


1) Batter piles are clearly effective in
restricting horizontal displacement of the
foundation in a liquefied soil flow environment.
The disadvantages of such piles are the larger
axial forces in the piles and the higher earth
pressures in the back of the footing (pile cap).
2) It is very important to strengthen quaywalls to prevent large displacements in soil, in
addition to other reinforcement measures of the
structures themselves. In any case, batter piles
deserve serious consideration as a means of
defending against large soil displacements.
REFERENCES
Tazoh, T., Ohtsuki, A., Fuchimoto, M., Nanjo, A.,
Yasuda, F., Fujii, Y., Nakahira, A., & Kuroda, C.
2002. Analysis of Seismic Damage to the Pile
Foundation of a Road Bridge Caused by the
Great Hanshin Earthquake, Shimizu Technical
Research Bulletin. (18), 1-25, Shimizu Corp.
Tazoh, T., Sato, M., & Mano, H. 2001. The Cause of
Ground Fissures Radiating from the Footing of a
Bridge Pier Generated by the 1995 Great
Hanshin Earthquake. Proceedings of Fourth
International Conference on Recent Advances in
Geotechnical Earthquake Engineering and Soil
Dynamics. Paper No. 4.57, 1-6.
Tazoh, T., & Gazetas, G. 1996. Pile Foundations
Subjected to Large Ground Deformations:
Lessons from Kobe and Research Needs.
Proceedings of the Eleventh World Conference
on Earthquake Engineering. Paper 2081.
Tazoh, T., Sato, M., & Gazetas, G. 2005. Centrifuge
Tests on Pile Foundation-Structure Systems
Affected by Liquefaction-Induced Soil Flow After
st
Quay Wall Failure, Proceedings of the 1
Greece-Japan Workshop on Seismic Design,
Observation and Retrofit of Foundation.
Tazoh, T., Shimizu, K., & Wakahara, T, 1987.
Seismic Observations and Analysis of Grouped
Piles, Dynamic Response of Pile Foundations,
Experiment,
Analysis
and
Observation,
Geotechnical Special Publication, No. 11, 1-20,
ASCE.

438

Table A-2: Installed sensors

Appendix
Table A-1 Materials and dimensions of the test
model

Laminar box
Pore water
Pile

Footing

Superstructure

Sheet
pile
quay-wall

Partition

Sensor
Accelerometer

Material, Size
Length: 805mm, Width: 475mm,
Height: 324mm (Inner size)
Silicon oil (30cSt) with 30 times the
viscosity of water
Material: Stainless steel
Number of piles: 4 (2 by 2)
Length: 270mm (prototype: 8.1m),
Outer diameter: 10mm (prototype:
30cm), Thickness: 0.2mm (prototype:
6mm)
Material: Steel
Thickness: 20mm (prototype: 60cm),
Length and width: 50mm by 50mm
(prototype: 1.5m by 1.5m)
Material: Steel
Rigid part: Length: 50mm, Width:
50mm, Height: 30mm (prototype: 1.5m,
1.5m, 0.9m)
Number of support columns: 4
columns, Thickness: 60mm, Size: 6mm
by 2mm (prototype: 18cm, 1.8cm by
0.6cm)
Material: Aluminum
A-side quay-wall (fixed to the base):
Height: 300mm, Length: 235mm,
Thickness: 10mm (prototype: 9m,
7.05m, 30cm)
B-side quay-wall (not fixed to the base):
Height: 250mm, Length: 235mm,
Thickness: 10mm (prototype: 7.5m,
7.05m, 30cm)
Material: Aluminum
Height: 305mm, Length: 750mm,
Thickness: 2mm (prototype: 9.15m,
22.5m, 6cm)

Pore
water
pressure
meter
Strain
gauge
Noncontact
displacem
ent meter
Earth
pressure
cell

439

Symbol
A

PP

Installed
locations
Shaking
table
Soil layer
Superstruc
ture
Footing
Quay-wall
Soil layer

BS

Pile

BS1 to BS4 (both sides A


and B)

Footing
Quay-wall

Disp-F (both sides A and


B)
Disp-Y (both sides A and
B)
EP (both sides A and B)

Disp

EP

Footing

Sensor notation
AG0
AG1 to AG4 (both sides A
and B)
AS1 (both sides A and B)
AF1 (both sides A and B)
Y1, Y2 (both sides A and
B)
PP1 to PP3 (both sides A
and B)

Why do pile-supported bridge piers and not abutments collapse in


liquefiable soils during earthquakes?
A. A. Kerciku1, S. Bhattacharya2 and H. J. Burd3
1

Balliol College (University of Oxford)


Departmental Lecturer in Engineering Science (University of Oxford)
3
University Lecturer in Engineering Science (University of Oxford)

Abstract
Failures of pile-supported bridges are still observed in liquefiable soils after most major earthquakes. In such

bridge failures, it is commonly observed that piers (intermediate supports) collapse, whilst the
abutments (end supports) remain stable. This paper investigates the reasons behind such occurrence.
Analytical investigations form the basis of this study. The pile-supported bridge structure has been
modelled as multi-frame system where the pile foundation forms the columns of the frame. The piles
are modelled as Fixed-Sway propped cantilevers where the propping action is provided by the
longitudinal stiffness of bridge deck. Previous research into the failure mechanisms of piled foundations
showed that buckling instability and bending are feasible failure modes. In contrast, this study investigates the
interaction between the bending and buckling failure modes. The importance of such interaction has been
highlighted by considering the failure of the well-documented collapse of the Showa Bridge during the 1964
Niigata earthquake.
INTRODUCTION

Roads and bridges are vital parts of the


infrastructure and therefore should remain in
working condition even after any natural
disaster such as a hurricane or an earthquake.
This is to facilitate the relief operations. Most
small to medium span bridges founded on
seismically liquefiable deposits (loose to
medium dense sands) are supported by pile
foundations. Failure of these pile foundations
has been observed in the aftermath of the
majority of recent strong earthquakes such as
the 1995 Kobe earthquake (JAPAN), the 1999
Kocheli earthquake (TURKEY) and the Bhuj
earthquake (INDIA). It has widely been
accepted that liquefaction-related effects are
the cause of these failures. It is a common
observation in liquefaction-related bridge failure
that piers (intermediate supports) collapse,
whilst the abutments (end supports) remain
stable, see for example Figs 1 and 2. Fig 1
shows the collapse of the Showa bridge after
the 1664 Niigata earthquake, while Fig 2 shows
the collapse of the Million Dollar bridge after the
1964 Alaska earthquake.

Fig 1: Collapse of the Showa Bridge, crossing the


Shinano River as a result of the 1964 Niigata
earthquake (JAPAN), After NISEE

The main aim of this study is to discuss why


it is that intermediate bridge supports seem
particularly prone to failure.

440

pile foundations supporting bridge piers are


predominantly designed to support the vertical
loads only.

Current understanding of pile failure

The most commonly adopted current


hypothesis of pile failure is based on a bending
mechanism. It is hypothesised that large inertial
and kinematic lateral loads produce bending
moments which exceed the capacity of the pile.
The inertial lateral loads are the result of the
earthquake induced inertial effects of the
superstructure and the kinematic loads are due
to flow of the soil following liquefaction and
strength degradation. The later effect, is
referred to by Krammer (1996) as lateral
spreading.

Fig 3: JRA (1996 or 2002) code of practice showing the


idealization for seismic design of bridge foundation.

Fig 4 presents the schematic diagram of a


typical two span bridge showing the abutment
and pier foundations. From static equilibrium,
for a multiple span bridge having similar span
lengths, abutments support a vertical load equal
to one-half of the vertical load supported by a
pier.

Fig 2: Collapse of the Million Dollar bridge, crossing the


former Copper River, following the 1964 Alaska
earthquake. After U.S. Geological survey data
series 1995

The Japanese Code of Practice (JRA 2002)


has incorporated this understanding of pile
failure and as shown in Fig 3. The code advises
practicing engineers to design piles against
bending failure assuming that the non-liquefied
crust offers passive earth pressure to the pile
and the liquefied soil offers 30% of the total
overburden pressure. Other codes such as the
USA code (NEHRP 2000) and Eurocode 8, part
5 (1998) also focus on the bending strength of
the pile. This simply treats piles as beam
elements and assumes that inertial lateral loads
and lateral spreading cause the bending failure
of the pile.

Fig 4: Schematic diagram of a typical two span bridge


showing the abutment and pier foundations.

Typically, the lateral load carrying capacity


of a pile is 10 to 20% of the axial load capacity.
Therefore, for a typical multiple span bridge
having similar span lengths, the number of piles
supporting an abutment is larger than the
number of piles supporting a pier.
As it can be observed from Fig 1 and 2,
collapse of pile-supported bridges in seismically
liquefied soils is often characterised by tilting or
failure of pier(s) and the subsequent collapse of
the bridge deck(s). It is worthwhile to note that
in these examples bridge piers collapsed whilst
the abutments remained stable.
Bhattacharya et al (2005) have shown that
the pile foundations supporting the Showa
Bridge piers (Fig 1) satisfy the JRA code (1996
or 2002) against bending due lateral spreading

Inconsistency between the current


understanding and the observed failure patterns

In bridge design, the number of piles


required to support an abutment is governed
partially by lateral load considerations since the
abutment has to retain earth and fills of the
approach roads; as well as vertical load
considerations of the deck. On the other hand

441

by a factor of approximately two. However,


these piles actually collapsed during the 1964
Niigata earthquake. Taking into account the
conservatism behind the JRA design method,
the collapse of the Showa Bridge suggests that
the failure of pile foundations may be influenced
by axial loads effects. This is in contrast to the
current design methods which concentrate on
lateral loads.
Buckling instability as an alternative explanation
Failure Explanation

Structurally, piles are slender columns with


lateral support from the surrounding soil. The
typical length to diameter ratio of most piles is in
the order of 25 to 100. When axially-loaded
piles lose lateral soil support due to liquefaction,
they behave like load bearing slender structural
members. Bhattacharya and Bolton (2005)
suggest that axially loaded piles may collapse
as a result of buckling instability due to the
removal of the soil bracing effects. This
hypothesis has been verified using dynamic
centrifuge testing and analysis of field case
records.
As the length of the pile increases, the
allowable load on the pile (from geotechnical
considerations) also increases due to the
additional shaft friction and the enhanced base
capacity offered at greater depths. On the other
hand, following Eulers formula, the buckling
load (if the pile were laterally unsupported by
soil) decreases inversely with the square of its
length. Fig 5 shows a typical plot for the
variation of allowable load P and buckling load
(if unsupported) Pc of a pile against length of
the pile. Therefore if a pile is unsupported
beyond a certain depth, buckling instability
becomes a possible failure mechanism.
Earthquakes impart lateral loads on structures.
Subsequently, the lateral loads lead to
considerable lateral displacements of piles in
liquefied soil, and to an increase of the effect of
out-of-line vertical forces. This is often termed
as the P- effect. For large displacements and
vertical loads, the P- effect can be a
significant contributor to the total bending
moment generated in pile foundations in
liquefied soil. Therefore, the analysis of the
interaction between the bending and buckling
failure modes are considered important in
understanding the behaviour of pile foundations
in liquefiable soils.

Fig 5: Allowable load and buckling load (if unsupported)


for a typical pile, Bhattacharya and Bolton (2004)

THE P- EFFECT

The P- can be described as a second


order geometrical effect which produces an
additional overturning moment. This effect is
discussed, for example in Wilson and
Habibullah (1987). For a cantilever column the
P- effect is governed by the axial load and the
lateral deflection of the tip.
To illustrate the P- effect, a cantilever
beam under the action of axial and lateral loads
is considered, as shown Fig 6. This might
represent a pile under the combined action of
axial and lateral loads, in the absence of soil
support. For illustration purposes, the cantilever
pile is modelled as a rigid bar-spring system
with the spring k representing the lateral
stiffness of the strut, where k is the lateral load
necessary to produce a unit tip displacement:

k=

3EI
L3

E is the piles Youngs modulus, I its second


moment of area and L the cantilevers length.
Two moment-equilibrium analyses are carried
out for the rigid bar-spring system:
a) First, a first order analysis, which omits the
geometrical and axial load effects (P-
effect).
b) Second, a second order analysis, which
incorporates the P- effect.

442

it is necessary to include the P- effect in the


bending moment calculations as well as any
concurrent moments produced by lateral
spreading or inertial loads.
ANALYTICAL SOLUTIONS

In this section we will formulate a simple


analysis, which investigates the conditions
necessary for failure at first yield of a pile
foundation in liquefied soil.
Fig 6: The P- effect: A cantilever can be modelled as a

Modelling the Piles

rigid bar-spring model. k is the structural stiffness


of the cantilever.

In a first order analysis, the tip displacement,


is assumed to be small and moment
equilibrium is independent of the axial load P.
By taking moments about the bottom hinge:

0 = FL (k) L
F = k

(1)

However, if a second order analysis is


applied, geometric change has to be taken into
account - by taking moments about the bottom
hinge and manipulating we can write:

F = k

P
P
= k
L
L

To gain an understanding of the behaviour


of pile foundations in liquefied soils the authors
have modelled the piles as beam-columns with
various boundary conditions. In practice, the
boundary condition of the pile depends on the
depth of embedment below the liquefied soil
and the moment restraint at the pile cap. For a
bridge foundation in liquefied soil the bridge
deck imposes a lateral restraint on the pile head.
It is assumed that the liquefied soil
possesses no lateral resistance, and that, as
suggested by Bhattacharya et al (2005), piles
are fixed in the non-liquefied strata at an
appropriate depth. The different boundary
conditions combinations can be identified.
These combinations often encountered i
practise are shown in Fig 7 (a), (b) and (c).

(2)

Eq 2 shows that, by including the P- effect the


overall stiffness of the system is reduced by P/L.
In a cantilever column the term P contributes
to the overall bending moments and is
significant only when the ratio /L is relatively
large. In structural engineering, the P- effect is
minimized by making the columns stiff or by
using cross bracing.
For piles in liquefiable soils, the /L ratio
can be considerable due to the following
reasons:
(a) Earthquakes induce lateral inertial or
kinematic loads which impart large
displacements to the superstructure and the
pile tips.
(b) Seismic liquefaction reduces the ability of
soil to provide lateral support to the pile.
Iwasaki (1984) suggests that large pile
deflections in the order of metres are observed
in sites where liquefaction occurred. Therefore,

443

Fig 7(a): Bridge pile model: Fixed-Free boundary


conditions, Due to the lack of moment resistance at the
pile cap, the pile tip is free to translate and rotate

depth in the non-liquefied soil layer, below the


liquefiable layer, and supporting structures
(bridges and buildings) is a fixed-sway
cantilever as described Fig 7(b) and (c).
The lateral and axial loads (inertial and
kinematic) acting on a bridge are shown in Fig 8.
qC(x) is the crust soil stratum passive earth
pressure, qL(x) is the distributed load due to the
soil lateral spreading and H , F and M are the
earthquake inertial vertical and horizontal
loading and applied moment. Further, P is the
permanent axial load on the piles due to the
bridge live and dead loads and the spring k
represents the lateral restraint imposed by the
bridge deck and the other piles supporting the
bridge. For simplicity, the combined effect of the
lateral loads may be modelled as the action of
one concentrated load F, as shown in Fig 8.

Fig 7(b): Bridge pile model: Fixed-Sway boundary


conditions. Due to moment resistance at the pile cap, the
pile tip is free to translate but not rotate

General analysis

In this section the interaction between the


bending and buckling failure modes is analysed.
The bending moment experienced by the pile is
dependent on the action of the lateral loading
and the P- effect, whilst the buckling stability
is a function of its geometrical and material
properties. In this section the governing
differential equations for the stability of the
system is formulated and solved. For the
purpose of this analysis the action of the
vertical and moment inertial loads has been
neglected. Fig 9 shows the fixed-sway
cantilever model considered in this section.

Fig 7(c): Bridge pile model: Fixed-Sway boundary


conditions when a non-liquefied crust is present. Due to
moment resistance provided by the non-liquefied crust, the
pile tip is free to translate but not rotate

For most bridges the connection between


the pile cap and piers is usually stiff and
moment resisting. This implies that the pile
head is free to translate but is unlikely to rotate
at the top. Therefore, the most appropriate
model for an end bearing pile fixed at some

Fig 8: Pile model: Fixed-Sway boundary conditions.

444

FL 2
L
tan 1

P L 2
=
kL 2k
L
1
tan
+
P P 2

(6)

To explain how the applied loads are related


to the pile failure we need to investigate the
load combination as the end reaction moment
M0 (from Eq 5) reaches a maximum allowable
value My, the bending moment at first yield. Eq
6 may be rewritten as:
Fig 9: Pile model: a Fixed-Sway cantilever

FL
P P
(7)
= 1 c + c
2M p
P P

tan

kL
2 EI
2
P
c ; P =

and =
where =
c
2
L
Pc
P
2 Pc

The equilibrium equation for the above


system is:

EI

d 2v
= Pv Fx + kx + M 0
d x2

(3)

where Mo is the end moment reaction, k is the


lateral stiffness experienced by the pile due to
the interaction with the rest of the structure and
EI is the flexural stiffness of the pile. x is the
depth below the soil surface, v is the
displacement measured as shown in Fig 9, and
is the maximum lateral displacement
(measured at the bottom of the pile). The
ordinary differential equation is solved in a
general form and the unique solution is found
by applying the boundary conditions for slope
and displacement at x=0:

v=

In Fig 10 the non-dimensional relationship


between FL/2Mp and P/Pc is plotted as a
function of increasing non-dimensional lateral
stiffness . For each specific curve, the area
confined by the curve and the axes represents
the safe design region, whilst the area outside
these borders represents the unsafe design
region.

M0
F
F k
M
x + 0 (4)
cos x +
sin x
P
P
P
P
where

= P EI .

By assuming system symmetry, the end


reaction moments at x=0 and x=L are
equivalent and equal to M0. Therefore, from
static equilibrium the end moment reaction can
be expressed as:

M0 =

1
[(F k )L + P ]
2

(5)
Fig 10: Non-dimensional design chart for different values
of k.

By applying the boundary conditions for slope


or displacement at x=L to Eq. (4):

445

As increases, the curve tends to cross the


P/Pc axis at 4.

In Fig 11(b) and * represent the lateral


stiffness (load required to produce a unit lateral
displacement of the head) of the pier piles and
the abutment piles respectively.

The bridge deck stiffness

This section aims to describe the lateral


restraint experienced by piles as a result of
interaction with the rest of the bridge structure.
This is necessary in order to complete
aforementioned model and respective design
framework with a methodology for estimating
the lateral stiffness k and corresponding nondimensional stiffness .
A pile supported bridge structure can be
modelled as multi-frame system as shown in
Fig 11(a), where each element Pi represents
the i-th pile from the left. The piles, modelled as
Fixed-Sway cantilevers, experience restraint to
lateral movement (in the longitudinal direction of
the bridge) due to the lateral stiffness imposed
by the other piles and the longitudinal stiffness
of bridge deck. To evaluate the stiffness
experienced by each pile it is necessary to
model the structure in Fig 11(a) as the springsystem described in Fig 11(b), where the
displacement of node i represents the tip
displacement of pile Pi, i.e. the pile head
displacement.

12 E p I

and * =

L3

12 E p * I *
L3 *

where E is the Youngs modulus of the pile, I is


the second moment of area, L is the distance
from the tip of the pile to its assumed fixity point
and E*, I*, and L* are the respective
dimensions of the abutment piles. Finally,
represents the longitudinal stiffness of the
bridge deck.

E d Ap
Ld

where Ap is the cross sectional area of the deck,


Ed is the decks Youngs modulus and Ld is the
length of the deck slab. The spring-stiffness
experienced by each pile Pi is the sum of the
stiffness due to the springs on the left kLi and
the stiffness due to the springs to the right kRi.
The magnitude of kLi and kRi is dependent on
the parallel and series sum of the individual ,
and * springs. Table 1 summarises the value
of kLi and kRi at various nodes i.
Table 1: Left and right stiffness of experienced by
the piles
Spring stiffness equation

kLi

kRi

1
2,3,4 n-2

1
k Li 1
1

n-1

Fig 11: The multi-frame structure in (a) represents a

+ k Ri +1
+

1
1

+ k Ri +1
+

k Li 1

1
1

1
k Rn

bridge in liquefied soil. This structure can be


modelled by the system in (b). The systems in (b)
and (c) are equivalent.

446

1
k Ln 1

Assuming that each deck span is identical and that the pierpiles also are all identical

the piles (as it might apply to abutments


designed mainly to serve as earth retaining
structures), the stiffness experienced by the
piles is dependent on its location along the
bridge, as shown in Fig 12. In this case the
central piles experience the lowest stiffness,
whilst the edge piers (P2 and Pn-1) experience a
stiffness n/4 times larger.

The total stiffness experienced by pile Pi is


the parallel sum of springs kLi and kRi as shown
in Fig 12(c), where kTi:

kTi = k Li + k Ri

(8)

From the equations in Table 1, it is possible


to evaluate the stiffness experienced by the
piles for some pile supported bridges. Some
typical combinations of deck, pier piles and
abutment piles stiffness, and the corresponding
restraint experienced by individual piles have
been summarised in Table 2.

THE FAILURE OF THE SHOWA BRIDGE

In this section the authors investigate the


failure of the Showa Bridge during the 1964
Niigata earthquake. A summary of the bridge
failure and current understanding is followed by
the authors critique of the assumptions made in
the failure hypothesis.
Finally, using the
analytical tools presented earlier, the authors
have attempted to explain the reasons why the
pier piles failed and the abutments remained
stable.
Background and current explanation of the
failure

The Niigata earthquake, with epicentre in


Niigata, Japan, occurred on the 14th of June
1964 and registered 7.5 on the righter scale.
Located some 55 km from the epicentre,
crossing the Shinano River, the Showa Bridge
was one of the three bridges which collapsed
completely as a result of the earthquake.

Fig 12: Variation of the lateral stiffness experienced by


the bridge piles, for

* >> and

/>10000.
It is interesting to note that when the
abutment stiffness is much larger than that of

Table 2: Stiffness experienced by the piles and applicability to the design method in Fig 10 for the various bridge design
combinations
Abutment pile
stiffness, *

Pier pile to deck


stiffness ratio, /

>10000

Equivalent lateral stiffness


experienced by pile Pi, kTi

E d Ap 1
1
+

Ld i 1 n i
for i=2,3,4, n-1

* >>
<0.0001

12 EI
L3

for all i

>10000

12nEI
L3

for all i

* =
<0.0001

12 EI
L3
for all i

447

Equivalent lateral non-dimensional


stiffness i

E d Ad L
2 Ld EI

1
1
+

i 1 n i

for i=2,3,4, n-1

0
for all i

12(n 1)

2
for all i

0
for all i

Fig 13: Showa Bridge detail: Soil Liquefaction profile. Adapted from Hamada and O'Rourke (1992). The soil behind the
left abutment is subjected to lateral spreading and creates a hinge near the abutment.

seismic inertial loading on the basis that


earthquake motion had stopped. Using aerial
photography Hamada and O'Rourke (1992)
studied the lateral spreading of liquefied soil
during the Niigata earthquake. They observe
that the riverbank to the left of the Showa
Bridge moved by 4 m towards the river centre.
Based on this information the authors
concluded that lateral spreading caused the
bending failure of the piles P5 and P6 and the
subsequent collapse of the bridge girders.

The collapse of the Showa Bridge has been


one of the central case histories in the study
liquefaction induced failure during the Niigata
earthquake. Most studies (Fukuoka (1966),
Hamada and O'Rourke (1992), and Yasuda and
Berrill (2000)), suggest that lateral spreading
was the direct cause of failure.
Reliable eyewitnesses quoted by Horii
(1966) suggest that the 303.9 m long and 24 m
wide superstructure collapsed 1-2 minutes after
the peak ground acceleration (PGA) had
ceased.
Evidence
suggests
that
the
phenomenon of delayed pile foundation failure
is typical for failure due to liquefaction related
effects. As showed in Fig 13 the sequential
collapse initiated when piers P5 and P6
collapsed in opposite directions. As a result
girder E fell in the river. Immediately afterwards,
in a domino effect, girders F, D, C and B
partially fell in the river. Hamada and O'Rourke
(1992) estimated the ground liquefaction profile.
As shown in Fig 14 the soil liquefied to a
maximum depth of 10 m below the riverbed and
to a maximum depth of approximately 5 m
below the riverbed near the left abutment.

Qualitative Argument

Whilst it is clear that lateral spreading


occurred around the left riverbank, there is no
evidence that the riverbed (directly underneath
the bridge) was significantly affected by lateral
spreading. In fact a potential energy argument
can prove the contrary: Liquefied soil will
attempt to flow towards a form of lower potential
energy. Therefore, liquefied soil on a slope will
flow downwards collecting at the end of the
slope, and liquefied soil behind a discontinuity
such as an abutment wall will tend to flow
across the discontinuity in order to reach a
lower potential energy form. As shown in Fig 14
the liquefied soil of the riverbed underneath the
Showa Bridge is already in its lowest energy
from and therefore will not spread laterally. It is
reasonable to believe that the lateral flow from
the left riverbank heaved close to left abutment,
and the soil near piles P5 and P6 remained
largely stationary. In support of this argument is
the fact that piers P5 and P6 collapsed in
opposite directions. Had lateral spreading
caused the bending failure of the piles, they
would have collapsed in the same direction.
Conclusively, it is reasonable to believe that the
flow of the lateral spreading of the soil on the
left bank of the river did not significantly affect
the collapsed the Showa Bridge pier piles.

Fig 14: Showa Bridge detail Adapted from Ishihara


(1984)

Based on the delayed collapse of the bridge,


Fukuoka (1966) and Hamada and O'Rourke
(1992) discard the possibility of collapse due to

448

The loads acting on the piles are:


a) Lateral load F (in the longitudinal bridge
direction), which is a result of the earthquake
induced permanent displacement
b) The bridge self weight P.

The structural stability of the pile


foundations was first put into question by
Bhattacharya and Bolton (2004). The authors
estimated the critical Euler buckling load
assuming the liquefied soil offer no lateral
restraint to a pile. It is shown that the factor of
safety against buckling instability of the piles
was close to one. Evidence of longitudinal
scratches provided by Horii (1966) suggests
that there was significant relative motion
between the girders and the pier-caps. These
displacements led to permanent geometrical
imperfections in the pile, and acted as a
catalyst to the P- effect.
In the context of a new failure hypothesis it
is reasonable to believe that the liquefaction
front travelling from top to bottom reached a
critical depth a few moments after the PGA. At
this critical depth, the combination of
superstructure loading, earthquake imposed
imperfections (in the form of pile tip
displacement), and reduced lateral support form
the liquefied soil, resulted in the collapse of the
Showa Bridge.

Finally, by assuming that the soil poses no


lateral restriction, piles P5 and AL can be
modelled as shown in Fig 15.

Fig 15: Showa Bridge pile model as a fixed sway


cantilever

Quantitative analysis

The analysis below aims to quantitatively


define the role that the bending and buckling
failure modes played in the 1964 collapse of the
Showa Bridge. More specifically the reason why
pier piles P5 failed and the left abutment piles AL
remained stable. According to Hamada and
O'Rourke (1992), as shown in Fig 14 the soil
underneath the Showa Bridge liquefied, to a
maximum depth of 10 m. Further to this Iwasaki
(1984) suggests that the bridge piers and
abutments were both supported on similar pile
bents (9 single-row piles of 609 mm diameter).
As shown in Fig 13 the Showa bridge deck
composed of panels, each alternatively resting
on roller and fixed supports. For the purpose of
this analysis we will assume the lateral loads
were large enough to overcome the roller
friction and therefore there was no lateral
restraint to the movement of the piles. In Table
2 this corresponds to * = and
/>10000. Therefore, each pile of the piers
and abutments pile bents can be analysed
separately using the modelling guidelines
suggested, and the design chart of Fig 10, for
=0.

As shown in Fig 16, Iwasaki (1984), the left


abutment moved by about 0.1 m towards the
centre of the river. Further, it can be estimated
from detail drawings in Iwasaki (1984) that the
maximum allowable pile crown displacement for
pier P5 during the earthquake, such that it would
not fall off the pier-cap, was 0.5 m. This
displacement information and the detailed
design data summarised by Bhattacharya et al
(2005) can be used to back-calculate the lateral
force that was necessary to cause the
displacements in question. By modifying Eq 5,
the lateral load necessary to cause the
observed displacement can be found by:

F=

Pc
tan
P
2

P
Pc

(15)

The magnitude of the axial load acting on the


abutment piles is approximately 30% that of the
axial loads acting on the piles. This is because
abutments support only one half of the 14 m
girder (as shown in Fig 13), whilst each pile
supports one 27 m long girder.

449

Fig 16: Deflections of the pile caps according to Iwasaki (1986)

Using this information we can now derive


P/Pc and FL/2My. To investigate the behaviour
of the pile with respect to yield failure we plot
the data points P/Pc and FL/2My in the design
chart of Fig 10. The computed numerical values
for the loads and corresponding moments are
presented in Table 3, and the non-dimensional
data points plotted in Fig17.
As shown in Fig 17 the design of the
abutment piles (as for the assumptions made) is
safe by a satisfactory margin, whilst the design
of the pile P5 is unsafe. The P- effect for P5 is
responsible for 20% of the total bending
moment. In the abutment piles, due to smaller
lateral displacements and axial loads, the P-
effect in is only responsible for 3% of the net
bending moment. Therefore, the contribution of
the P- effect is 6 times larger on P5 than on AL.
Further, the axial load in the pier-pile P5 is
approximately 3.5 times larger than the axial
load in the abutment-pile AL, and the P5 lateral
load in is approximately twice the AL lateral load.
This suggests that the contribution of the P-
effect to the net bending moment increases
rapidly with P and F.
Had the pier pile P5 head moved only by 0.4
m the design would still have proven to be
unsafe (FL/2Mp=0.274, P/Pc=0.197). The
unsatisfactory design performance of the pier
piles compared to the abutment piles is due to
the larger axial loads acting on the piers the
result of supporting a larger portion of the girder.
This emphasizes the important role that axial
loads play in the performance of piled

450

foundations in liquefied soils. As a final remark


it is interesting to note that, as shown in Fig 17,
pile P5 is designed safely against buckling or
bending taken as isolated failure modes.
However, it is an unsafe design when buckling
and bending are considered as interactive
failure mechanisms.
Table 3: Bending moment and buckling calculations
for pile P6 and the abutment piles
Numerical values
Variable

Pile P5

Left abutment
piles

(m)

0.50

0.10

P (N)

750000

205000

EI (Nm )

16000000
0

160000000

L (m)

21

17

F (N)

82178

37632

FL (Nm)

1725737

639749

My (Nm)

629000

629000

Pc (N)

895000

1366035

P/Pc

0.838

0.150

References
From Fig 16,
according to
Iwasaki (1986)
Bhattacharya
(2005)
Hamada and
O'Rourke
(1992)
Bhattacharya
(2005)
Eq (20)
Bhattacharya
et al (2005)
Bhattacharya
et al (2005)

0.533
0.197
FL/2Mp
taken as 35% of the 750000 N load acting on the piers, which
would correspond to the weight that half of the 13m girder
imposed on each pile.
2
Estimated from Fig. 8 p3-21, depth of fixity taken as 6D, where D
is the pile diameter.
3
My=I/y for =500MPa
4
2
2
Buckling load for a Fixed-Sway cantilever, Pc= EI/L ,
calculated using data in Bhattacharya et al (2005).
1

lateral restraint offered by the bridge deck to


the bridge foundation has been proposed.
For certain bridge configuration, the restraint
offered by the deck is lowest in the middle of
the bridge.
4. Lateral spreading is not of the major
contributor to the failure of the Showa
Bridge. It is reasonable to believe that the
flow of the lateral spreading of the soil on
the left bank of the river did not significantly
affect the collapsed piles of the Showa
Bridge. The combination of superstructure
loading, initial imperfections and reduced
lateral support form the liquefied soil
resulted in the collapse of the Showa Bridge.
The Showa Bridge failed due to lack of
structural stability of the piled foundations.

Fig 17: Design chart for piles in liquefied soils, modelled


as the Fixed-Sway cantilever of Fig 13. The data
points for the design performance of the
abutment pile AL and pier pile P5.

CONCLUSIONS
1. The axial loads can be one of the major

contributors to the collapse of pile


foundations in liquefied deposits. The axial
loads contribute directly to the bending
moment through the P- effect, and also act
as an important destabilising factor
indirectly affecting the magnitude of . It
was demonstrated by re-analysing the
performance of the abutment piles and the
pier piles of the Showa Bridge that the
contribution of the P- effect to the net
bending moment increases rapidly with the
axial and lateral loads. Therefore, the P-
effect should be included in the bending
moment calculations while analysing the
effects of lateral spreading and/or inertial
loads. It has been shown that P- is more
pronounced in bridge piers. This may
explain the observation that bridge piers
collapse while the abutments remain stable.

REFERENCES
Bartlett, S. F. and Youd, T. L. (1992) Empirical
Analysis of Horizontal Ground Displacement
Generated by Liquefaction Induced Lateral
Spreads Tech. Rept. NCEER 92-0021, National
Center for Earthquake Engineering Research,
Buffalo, NY, USA.
Yasuda S. and Berrill, J.B. (2000) Observations of
the earthquake response of foundations in soil
profiles containing saturated sands, Proc Geo
Eng2000 Conf. Melbourne,,
Bhattacharya S. and Bolton M. D. (2004). Errors in
design leading to pile failures during seismic
liquefaction. Proceedings of the 5th International
conference on case histories in earthquake
engineering, New York, USA
Bhattacharya S., Bolton M. D. and Madabhushi S. P.
G. (2005). A reconsideration of the safety of piled
bridge foundations in liquefied soils. Soils and
Foundations. Vol. 45, No. 4, p13-24.

2. Designers should investigate the interaction


of the failure due to the bending and
buckling modes. The study shows that a
piled foundation in liquefied soil may be
designed safely against the buckling and the
bending modes but can be unsafe when the
two failure modes may interact.

Eurocode 8: Part 5 [1998]. Design provisions for


earthquake resistance of structures- foundations,
retaining structures and geotechnical aspects,
European Committee for standardization, Brussels.
Fukuoka M., (1966). Damage to civil engineering
structures. Soils and Foundations, Vol. VI, No. 2,
p45-52.

3. The restraint offered by bridge decks to the


bridge foundations (abutments and piers)
may vary depending on the configuration of
the bridge, relative stiffness of the piles and
the deck. A methodology to estimate the

Hamada M., and O'Rourke T. D. (1992). Case


studies of liquefaction and lifeline performance
during past earthquakes Volume 1: Japanese

451

case studies. Technical Report NCEER-92-0001,


National Centre for Earthquake Engineering
Research, Buffalo, NY, USA.

Kawakami F. and Asada A. (1966). Damage to


ground and earth structures by the Niigata
earthquake of June 16, 1964. Soils Foundations,
Vol. VI, No. 1a, p14-30.

Horii K. (1968). General report on the Niigata


earthquake. Tokyo Electrical Engineering College
Press. Part 3: Highway Bridges, p431-450.

Krammer, S. L. (1996) Geotechnical Earthquake


Engineering, Prentice Hall New Jersey

Horne M. r. and merchant W. (1965). The stability of


frames. Pergamon Press, Oxford, UK

NISEE: National Information Services for earthquake


engineering, University of California Berkley

Iwasaki T. (1984). A case history of bridge


performance during earthquakes in Japan. Int'l
Conf. Case Histories in Geotechnical Engineering.
University of Missouri-Rolla, Vol. 3, p981-1008.

U.S. Geological survey data series, DDS-21, 1995.


Earth Science photographs from U.S. Geological
survey Library
Wilson E. L. and Habibullah A. (1987) Static and
Dynamic Analysis of Multi-story Buildings Including
P-Delta Effects, Earthquake Spectra, Earthquake
Engineering Research Institute, Vol. 3, No. 3

Iwasaki T. (1986). Soil liquefaction studies in Japan,


State-of-the-art. Technical Memorandum No. 2239,
Public Works Research Institute, Tsukuba, Japan,.

JRA [1996]. Japanese Road Association,


Specification for Highway Bridges, Part V, Seismic
Design.

452

Effect of excess pore water pressure buildup on building damage


P. Dakoulas1
1

University of Thessaly, Volos, Greece

Abstract
The effect of the presence of a liquefiable soil layer on the seismic behaviour of a typical
actual 7-storey reinforced-concrete building is investigated. Four identical buildings,
designed on the basis of the low ductility requirements of earlier seismic codes, are
assumed to be founded at four actual sites. The first site represents a typical soil profile with
average values of soil properties within a seismic zone. Located in the same zone, the other
three sites contain a liquefiable soil layer, with relative densities 60%, 50% and 40%,
respectively. The building at the typical soil profile experienced tolerable deformations.
However, the three buildings founded at the sites with the liquefiable layer developed
significant interstorey drift, and plastic deformation and bending in many beams, especially
between the ground floor and the third floor.

specific sites in the same seismic zone, each


containing a liquefiable layer of different relative
density. These four soil-structure systems are
subjected to a series of seismic excitations that
have the characteristics of the design
earthquake for that region. The response and
the seismic damage of the four systems are
examined in order to assess the effect of the
presence of the liquefiable layer.
Before proceeding with the analysis of the
soil structure system, it is of interest to
examine first the site seismic response

INTRODUCTION

The potential for significant seismic damage


of reinforced-concrete buildings, designed on
the basis of the low ductility requirements of
earlier seismic codes, is of great interest as it
affects the safety of a large proportion of
buildings worldwide. Buildings designed using
the Greek seismic code of 1985 (or earlier) may
have a natural ductility 2.5 and, therefore,
posses a limited capacity to sustain large
deformations. If founded on soils that contain a
liquefiable layer, such buildings may experience
significant seismic damage. The objective of
this study is to examine the effects of the
presence of a liquefiable foundation soil layer
on the seismic damage of such buildings.
The study examines the seismic behavior of
a typical 7-storey building that has been
designed according to the outdated Greek
Seismic Code of 1985. The same building is
assumed to be founded on four different soil
profiles located at the center of the city of
Larissa, Greece. The first building is founded on
a typical soil profile, i.e. having average shear
wave velocities, of a certain seismic zone at the
city center. The other three buildings are
assumed to be founded on soil profiles at

SITE RESPONSE

A seismic hazard analysis study for the city


(Pitilakis and Tsotsos 1995) has proposed two
earthquake scenarios, one for a near field
source and one for a far field source. The
characteristics of the two seismic scenarios are
given in Table 1. The aforementioned study
divides the city in 18 seismic zones. This paper
focuses only on Zone 6, located at the city
center, near the banks of Pinios River (Fig. 1).
The typical soil profile of this zone is subjected
to ten seismic excitations, utilizing the historic
records given in Table 2, properly scaled to the
expected peak ground acceleration.

453

Pinios River

Zone 6

Fig 1: Seismic zone 6 at the center of the City of Larissa

(a)

Zone 6

0
(b)

Vs, m/s

500

1000

Fill
- Debris

20

Clay

30
40

20

Clay

-
Silty
Sand
- Silt

- -Silt

Silty
Sand

Depth , m

Depth , m

(m)

10

40
60
80

Clay

100

100

Fig 2: (a) Soil profile and (b) S-wave velocity vs depth at Zone 6 (Pitilakis & Tsotsos 1995)

greater than -100 m, the material is assumed


to be equivalent to bedrock having a shear
wave velocity VS = 800 m/s. The initial

Fig. 2(a) shows the typical soil profile at Zone


6. The soil profile consists of a 10 m thick
layer of fill, a 3 m thick layer of clay, a 4 m
thick layer of silty sand and silt, a 10 m thick
layer of clay, a 10 m thick layer of silty sand
and silt, and a very deep layer of clay. Fig.
2(b) plots the distribution of the shear wave
velocity versus depth at Zone 6. At depths

fundamental period of the site is T1 = 0.88


seconds.
The site response is evaluated using 1-D
equivalent linear analysis. The variation of the
shear modulus ratio and the critical damping

454

considering an equivalent frame section along


its shorter side.
The frame consists of 5 columns, as
shown in Fig. 4(b). The beams and columns
of the building are modelled as elastic-plastic
members. Fig. 4(c) plots representative
relationships between the ultimate moment
and the axial force at various floor levels for
columns K2 and K4. It is noted that the
concrete strength reduction due to cracking
and the shear failure are ignored.

ratio with the amplitude of cyclic shear strain


is based on laboratory data by Pitilakis and
Tsotsos (1995). Figs. 3(a) and (b) plot the
distribution with depth of peak accelerations
and peak shear strains, respectively,
experienced during shaking by the ten
excitations listed in Table 2. The peak
accelerations at the ground surface range
from 0.30 g to 0.44 g.
Table 1: Seismic scenarios

B. Constitutive model for cohesionless soil


Seismic Scenario

Earthquake Magnitude, Ms

6.3

Epicentral Distance, R

80 km

6 10 km

Peak Ground Acceleration

0.20 g

0.335 g

Effective stress analysis is utilized for the


simulation of the development of excess pore
water pressures based on full coupling of the
volumetric changes of the soil skeleton and
the pore water. The constitutive model for
cohesionless soil used in the present study is
the model developed by Pastor et al. (1990),
subject to minor modifications. The model is
based on the generalized plasticity theory and
can be used for the monotonic and cyclic
behavior of cohesionless soil. It utilizes a nonassociated flow rule. Detailed descriptions of
the basic model are given by Pastor et al.
(1990) and Zienkiewicz et al. (1999).
For loose contractive sand the model
predicts the densification and strain hardening
under drained shearing, and the development
of excess pore pressure and liquefaction
under undrained shearing. For very dense
dilative sands in drained shear, the model
accounts for strain softening and residual
conditions at the critical state. A systematic
series of comparisons between predictions of
the modified model and experimental data of
different sands from both monotonic and
cyclic tests in compression-extension and
simple shear showed very good agreement
(Dakoulas 2003). As a representative
example, Fig. 5 compares the Cyclic Stress
Ratio (CSR) for initiation of liquefaction in
simple shear tests from model predictions and
experimental data of Nevada Sand (Arulmoli
et al. 1992) and Monterey Sand (DeAlba et al.
1976, as modified by Seed and Harder 1990)
for relative densities of 40 % and 60 %. The
model
seems
capable
of describing
realistically soil behavior under monotonic and
cyclic loading for a wide range of relative
densities. It is being used here to simulate
approximately the soil response under seismic
conditions.

Table 2: Earthquake excitations


Number

Earthquake

Year

Scenario

Almyros

1987

Edessa No. 1

1990

Edessa No. 2

1990

Kilkis

1990

Kobe

1995

Aegio

1995

El Centro

1940

Ierissos

1983

Kalamata

1986

10

Ouranoupolis

1983

SOIL - BUILDING RESPONSE


A. Model discretization

The seismic analysis is performed using


the program FLAC (Itasca 2000) and an
appropriate elasto-plastic model for the cyclic
behaviour of cohesionless soils. A typical
discretization of the soil profile structure
system is portrayed in Fig. 4(a), in which the
thickness of the individual soil layers may vary
to represent the soil profile at a given site and
is taken less than min / 8 , where min is the
minimum wave length of the excitation. The
building consists of a basement of 4 m depth,
a ground floor of 6 m height, and six additional
floors of 3 m height each. The foundation
consists of a mat with connecting beams. The
building is discretized as a 2-D structure by

455

(a)

Depth, m

-20

-40

-60

-80

-100
0.1

0.2

0.3

0.4

0.5

Acceleration, g

(b)

Depth, m

-20

-40
Aegio
Almyros
Kobe
Edessa 1
Edessa 2
El Centro
Ierissos
Kalamata
Kilkis
Ouranoupolis

-60

-80

-100
0.00

0.05

0.10

0.15

0.20

0.25

Maximum Shear Strain, %


Fig 3: (a) Peak acceleration and (b) Peak shear strain vs depth at Zone 6

456

0.30

0.35

(a)
Monterey Sand
Nevada Sand
Model

0m

Dr = 40%

110 m
100 m
(b)

12

Base
Gr. Fl.
Floor 1
Floor 2
Floor 3
Floor 4
Floor 5
Floor 6

10
8

Axial Force, MN

Dr = 60%

Fig 5: Cyclic Stress Ratio for initial liquefaction in


simple shear tests from model predictions
and experimental data for Nevada Sand
(Arulmoli et al. 1992) and Monterey Sand
(DeAlba et al. 1976, as modified by Seed
and Harder 1990): (a) Dr = 40% and (b) Dr =
60%.

4
2
0
-2

(c)

C. Building response at non-liquefiable profile


of Zone 6

-4
0.0

0.2

0.4

0.6

0.8

In the first case, the building is assumed to


be founded on the typical soil profile of Zone 6
(see Fig. 2), in which no liquefaction or soil
failure in general is expected. First, a static
analysis is performed for the evaluation of the
initial stresses. For the dynamic analysis of

1.0

Moment, MN.m

Fig

4: (a) Soil and structure discretization


(b) equivalent frame section for 2-D analysis
(c) relation of ultimate moment and axial force
at various floors for columns K2 and K4.

457

ground floor. Instead, huge glass windows are


placed, as they are preferred by the shop
owners at that floor. Hence, the building is
simulated as almost fixed at the basement,
and relatively more flexible than it is in reality
from the first floor to the top. It is expected
that, being more flexible, the ground floor may
experience a little higher deformation than the
computed in this analysis.

the soil structure system, the elastic moduli


and the critical damping ratios of the nonliquefiable soil layers are obtained from
equivalent linear analysis, so that they are
compatible with the amplitude of the seismic
shear strains developing in each material
element.
The radiation of wave energy back into the
bedrock material is accounted for by applying
the excitation in terms of seismic shear
stresses at the base of the soil-structure
system and using appropriate dashpots. The
free field response at the left and right side of
the soil-structure mesh are simulated using
two soil columns that are connected to the
mesh with dashpots.
Figures 6(a) and 6(b) plot the maximum
and residual horizontal relative displacements
along the height of the building for the ten
earthquake excitations. The maximum value
of the horizontal relative displacement is 12
cm, whereas the maximum value of the
horizontal residual displacement is about 4
cm. Fig. 7 plots the displacement time history
at various floor levels of the building subjected
to
the
Aegio
seismic
record.
The
corresponding accelerations at the top of the
building vary from 0.5g to 0.8g, whereas at
the base vary from 0.2g to 0.5g.
Figure 8 plots the distribution of the
Interstorey Drift Index (IDI), which is defined
as the ratio of the maximum relative
displacement between the ceiling and the
base of a given floor over the height of that
floor. Considering that the building has been
designed with the natural ductility ( 2.5),
the value of IDI should not exceed the value
of 1% (Bachmann 1998). The values of IDI in
Fig. 8 remained significantly less than 1%
and, therefore, the associated deformations
are tolerable. Also, the maximum values of
IDI occurred between the ground floor and the
third floor. A detailed examination of the
behavior of the column and element
connections
has
shown
that
plastic
deformations developed in most of the
connections of the superstructure. In this way,
the dissipated seismic energy was distributed
in a large portion of the building.
A limitation of the analysis is that it ignores
the stiffness of the non-bearing walls. Such
walls do not exist along the perimeter of

D. Building response at sites A, B and C

Fig. 9 portrays the soil profiles at sites A, B


and C, which are located within Zone 6 and
each contains a liquefiable layer. Table 3
gives the depth of the liquefiable layer, the soil
type, the field N-SPT blow counts, and the
estimated relative density for each of the three
layers. Simplified analysis of the liquefaction
potential using the NCEER Workshop method
and the Japan Road Association method,
which account for the influence of fines
content and plasticity index, has shown that
the factor of safety (FS) against liquefaction
for the design earthquake is less than one.
These values of FS are given also in Table 3
for the three sites.
Initially, the free field response at each site
is computed, followed by the analysis of the
soil structure system. The soil structure
system mesh that is used for the analysis is
similar to that in Fig. 4(a), with minor
modifications in the thickness of specific soil
layers. The liquefiable soil layers and those
that may experience soil failure are modelled
using the elasto-plastic constitutive model.
The model parameters are computed based
on the relative density of the soil, estimated by
using the corrected N-SPT values and the
grain size distribution characteristics. For
computational efficiency, the non-liquefiable
and non-failing (deeper) soil layers are
modelled as linear elastic materials having
strain-compatible dynamic properties derived
from previous equivalent linear analysis.
For the soil-structure system at site A,
subjected to the Kalamata earthquake record
scaled at 0.33g, Fig. 10(a) plots the
distribution of the excess pore water pressure
ratio u / m 0 within the soil layer at time t =30
s ( u = the excess pore pressure and m 0 =
the mean initial effective stress).

458

(a)
25

Height, m

20

15
Aegio
Almyros
Edessa 1
Edessa 2
El Centro
Ierissos
Kalamata
Kilkis
Kobe
Ouranoupolis

10

0
0

10

12

14

Maximum Displacement, m

(b)
25

Height, m

20

15
Aegio
Almyros
Edessa 1
Edessa 2
El Centro
Ierissos
Kalamata
Kilkis
Kobe
Ouranoupolis

10

0
0

Residual Displacement, m
Fig 6: (a) Maximum relative displacements and (b) residual relative displacements along the height of the
building at a site with typical properties of Zone 6 for 10 earthquake excitations.

459

Displacement, cm

2

4
2

10

12

14

Time, s
Fig 7: Displacement time histories at various floor levels for the Aegio earthquake excitation

25

Aegio
Almyros
Edessa 1
Edessa 2
El Centro
Ierissos
Kalamata
Kilkis
Kobe
Ouranoupolis

Height, m

20

15

10

0
0.00

0.20

0.40

0.60

0.80

1.00

Interstorey Drift Index, %


Fig 8: Interstorey Drift Index along the height of the building at a site with typical properties of Zone 6

460

Site A

Site B

DEBRIS
CL
SP-SW
CH

W.T.

Site C
DEBRIS
CL
SP-SW

DEBRIS
W.T.

SM-ML

CH

SW
CH
SM-SC
CH
SM-SC
CL
CH

SP-SW

CL

CL

CL

Fig 9: Soil profiles at sites A, B and C.

again significant plastic deformation of beams.


The deformed shapes of the building shown in
Figs. 10, 11 and 12 (at time t = 30 s), do not
represent the final stage, as additional
deformation may occur during excess pore
water pressure dissipation and consolidation
of the soil layers.

Fig. 10(b) plots the evolution of u / m 0 at


point M, located in the middle of the
liquefiable layer, beneath the central column
of the building. Fig. 10(c) plots the evolution of
relative displacements at various floor levels.
Similar results are presented for sites B
and C in Figs. 11 and 12, respectively. It is
evident from Figs 10 to 12 that the excess
pore pressures increase with decreasing
relative density. The residual values of
u / m 0 at point M for sites A, B, and C are
0.3, 0.4 and 1. At sites A and B, higher values
of the ratio occur near the sides of the
building, due to rocking motion. At site C,
liquefaction occurs quickly (within 5 seconds
from the start of the shaking) and expands
across the entire width and thickness of the
soil layer.
Fig. 13 plots the residual displacements at
all floors for the three buildings subjected the
Aegio, Kalamata and El Centro excitation
records.
For the same excitations, Fig. 14 plots the
IDI versus the building height. Note that
although liquefaction occurs at site C, the
damage at sites B and A (at time t = 30 s)
appears to be higher. This may be misleading:
the softening of the soil layer very early during
shaking at site C acts as a filtering
mechanism leading to smaller horizontal
displacements, but not necessarily smaller
overall damage, as the increased differential
movements among various columns cause

Table 3. Characteristics and factors of safety for


the three liquefiable layers
A

Depth of layer, m

4.6 6 m

4 6.2 m

59m

Soil type

SP - SW

SP - SW

SM - ML

N SPT

12

60 %

50 %

0.83

0.62

0.81

0.57

Site

Relative Density, Dr
Factor of Safety
(NCEER)
Factor of Safety
(JRA)

461

3
40 %
0.71

0.63

(a)
Site A

u / m 0
M

(b)

Excess pore pressure ratio

1
0.8
0.6
0.4
0.2
0
0.2

10

10

15

20

25

30

15

20

25

30

20

(c)

Displacement: cm

10

10

20

Time: s

Fig 10: Site A: (a) Contours of excess pore pressure ratio within the liquefiable layer and deformed structure
caused by Kalamata earthquake (time t = 30 s) (b) Excess pore pressure ratio time history at point M and
(c) Displacement time histories at various building levels.

462

(a)

Site B

u / m 0

Excess pore pressure ratio

(b)
0.8
0.6
0.4
0.2
0
0.2

10

10

15

20

25

30

15

20

25

30

20

(c)

Displacement: cm

10

10

20

Time: s

Fig 11: Site B: (a) Contours of excess pore pressure ratio within the liquefiable layer and deformed structure
caused by Kalamata earthquake (time t = 30 s) (b) Excess pore pressure ratio time history at point M and
(c) Displacement time histories at various building levels.

463

(a)
Site C

u / m 0

(b)

Excess pore pressure ratio

1
0.8
0.6
0.4
0.2
0

0.2

10

10

15

20

25

30

10

(c)

Displacement: cm

5

10

15
15

20

25

30

Time: s
Fig 12: Site C: (a) Contours of excess pore pressure ratio within the liquefiable layer and deformed structure
caused by Kalamata earthquake (time t = 30 s) (b) Excess pore pressure ratio time history at point M and
(c) Displacement time histories at various building levels.

464

Aegio

Site A

25

Site A

25

Kalamata
El Centro

20

Height: m

Height: m

20

15

10

15

10

Aegio
5

Kalamata
El Centro
2

10

12

0.5

IDI: %

Displacement: cm

Aegio

Site B

25

Site B

25

1.5

Kalamata
El Centro
20

Height: m

Height: m

20

15

10

15

10

Aegio
5

Kalamata
El Centro
2.5

7.5

10

12.5

15

17.5

0.5

20

Displacement: cm

1.5

IDI: %
Site C

25

Site C

25

2.5

Aegio
Kalamata
El Centro

20

Height: m

Height: m

20

15

10

15

10

Aegio
5

Kalamata
El Centro
1

0.25

Displacement: cm

0.5

0.75

IDI: %

Fig 13: Residual displacements along the height of


the building for sites A, B and C

Fig 14: Interstorey Drift Index along the height of the


building for sites A, B and C

465

CONCLUSIONS

REFERENCES

The effective stress method and an elastoplastic constitutive soil model have been used
to examine the effect of a liquefiable foundation
soil layer on seismic damage of a typical 7storey building. The parameters of the
constitutive model have been derived based on
the relative density of the soil, estimated from
corrected N-SPT values and the grain size
distribution characteristics. The results of the
seismic analysis of four different soil-structure
systems showed that the presence of a
liquefiable layer may affect significantly the
seismic behaviour and the damage of a
structure. The typical, low-ductility 7-storey
building investigated, experienced tolerable
deformations at the typical soil profile, in which
average soil properties are considered. In this
case the Interstorey Drift Index (IDI) remained
well below the critical values of 1%. At three
sites containing liquefiable soil layers with
relative density 60%, 50% and 40%,
respectively, the same 7-storey building
developed significant damage and values of IDI
well above 1%. Such deformations cannot be
sustained by the low natural ductility of the
older buildings designed by outdated older
codes, such as the Greek Seismic Code of
1985. In addition to horizontal displacements,
significant plastic deformation and bending was
found in many beams located between the
ground floor and the third floor. It is noted that
the conclusions regarding the behaviour of the
superstructure are drawn from the analysis of a
single typical building founded at four different
sites. It is desirable to examine the behaviour of
other typical structures and the effect of the 3-D
geometry of the structure on the seismic
behaviour and damage. Finally, it is noted that
the dynamic SSI interaction phenomena under
the influence of excess pore water pressure
and potential liquefaction are very complex and
affected by several factors. The effective stress
method can be a valuable tool in investigating
the effect of such factors.

Technical Chamber of Greece (2000),


Utilization of the Microzonation Study of the
City of Larissa, Larissa, (in Greek).
Arulmoli, K., Muraleetharan, M., Hossain, M
and Fruth, L., (1992), Verification of
Liquefaction Analyses by Centrifuge Studies
Laboratory Testing Program Soil Data
Report, Report, The Earth Technology
Corporation, Long Beach, California.
Bachmann, H., (1998) Seismic Protection of
Structures, Gourdas, Athens, (in Greek).
Castro, G. (1969), Liquefaction of sands,
Ph.D. Thesis, Harvard University, Harvard
Soil Mechanics Series, No. 81.
Dakoulas, P. (2003), Verification of a
constitutive model for non-cohesive soils,
Research Report, Univ. of Thessaly, Volos,
Greece, (in Greek).
Dakoulas, P. and Gazetas, G. (2005), Seismic
Effective Stress Analysis of Caisson Quay
Walls: Application to Kobe, Soils and
Foundations, Vol. 45(4), 113-125.
DeAlba, P., Seed, H. B., & Chan, C. K. (1976),
Sand liquefaction in large scale simple
shear tests, Journal of the Geotechnical
Engineering Division, ASCE, Vol. 102, 9, pp
909-927.
2000 (Greek Code for Seismic Design
2000), O.A.S.P. & Society of Civil Engineers
of Greece, Athens.
Pastor, M., O. Zienkiewicz, O., & Chan, C. H.,
(1990 ) Generalized plasticity and the
modeling of soil behavior, International J.
of Numerical and Analytical Methods in
Geomechanics, Vol. 14, pp. 151-190.
Pitilakis, K. and Tsotos, S. (1995),
Microzonation Study of the City of Larissa,
Aristotle University of Thessaloniki (in
Greek).
Itasca (2000), FLAC, Computer Software,
Manuals version 4.
Schnabel, P, Lysmer, J, and Seed, H.B.,
(1972) SHAKE: Equiv. Linear Seismic
Response Analysis of Horiz. Layered Soil
Deposits, Computer Software Manual.

466

Numerical Analysis of Gravel Drain Performance in Liquefiable Soils


A. G. Papadimitriou1, M.-E. Moutsopoulou2 , G. D. Bouckovalas3
1

Lecturer, University of Thessaly, Greece


M.Sc. Student, University of California at Berkeley, USA
3
Professor, National Technical University of Athens, Greece
2

Abstract
This paper studies the use of gravel drains as a means of mitigating earthquake-induced
liquefaction in non cohesive soils. The study is performed numerically using the 2D finite
difference code FLAC and a recently proposed bounding surface plasticity model for non
cohesive soil behavior under cyclic loading. The reliability of the numerical methodology is
verified via the simulation of a centrifuge test for a uniform 19m-thick liquefiable sand layer
that was improved with gravel drains and subjected to earthquake motion. The comparison
of data to simulations is performed in terms of time histories of accelerations and excess
pore water pressure ratios at various locations within the sand layer and is found
satisfactory. Given the accuracy of the methodology, parametric analyses were performed
for the study of the merely horizontal dissipation of excess pore pressures provided by
gravel drains in a thin liquefiable sand layer enclosed in practically impermeable clay layers.
The emphasis of the analyses is the rate of excess pore pressure buildup in the improved
ground, as compared to the recommendations of available methods for design of gravel
drains. In particular, the conservatism of the design charts of Seed & Booker (1977) is
discussed and preliminary recommendations are provided for their future use.

introduced in the methodology via an


empirical fitting of related data from a large
number of undrained element tests with
constant cyclic stress-ratio CSR.
Since its proposal the foregoing method has
been studied by various scientists aiming at the
amelioration of its predictions. Of these newer
method variants, the one that has attracted the
most attention is the work of Onoue (1988),
who retained assumption (c) as is, but used the
drain permeability kd as a design parameter that
creates the so-called drain resistance L, which
reduces the dissipating capability of drains if the
kd value is not significantly larger than that of
the natural soil ks. Furthermore, Onoue (1988)
studied the problem of combined horizontal
(towards the drains) and vertical upward
dissipation of excess pore pressures through
the sand layer to the ground surface that may
occur if the layer in question is not covered by a
low-permeability layer (e.g. a clay cap). In
parallel, Matsubara et al (1988) and Iai and
Koizumi (1986) proposed similar method

INTRODUCTION

The construction of gravel drains and/or


gravel piles is a commonly used method for
mitigating earthquake-induced liquefaction in
non cohesive soils. Among the advantages of
this improvement method is the fact that the
construction method is relatively simple and its
cost is relatively small. The relevant design
methods are based on the pioneering work of
Seed & Booker (1977), which assumes that:
(a) The dissipation of excess pore pressures is
based on purely horizontal axisymmetric
flow towards the drains
(b) The drain material does not increase the
overall stiffness of the improved ground (i.e.
Gd = Gs) and has practically infinite
permeability (i.e. kd , the exact value of
which has no effect on the dissipated
excess pore pressures, if it is at least 200
times larger than the permeability of the
natural sand, ks)
(c) The rate of excess pore pressure buildup is

467

variants, with relatively small differences on the


dissipating capacity of the drains, as a function
of the drain resistance L factor. More recently
Han & Ye (2001) proposed the use of modified
consolidation coefficients cv and cr for the
vertical and the horizontal dissipation of excess
pore pressures, as a function of the relative
compressibility of the drain mv,d and the natural
soil mv,s materials, but this in connection to the
acceleration of consolidation of clays.
Based on the above, it may be concluded
that the design methods of gravel drains have
attracted a lot of attention over the last 30
years. In particular, the relative importance of
assumptions (a) and (b) has been thoroughly
studied
over
the
years,
but
these
methodologies cannot be considered other than
method variants of the work of Seed & Booker
(1977), since assumption (c) has remained the
underlying foundation of most of these works.
Furthermore, in practice, the original design
charts of Seed & Booker (1977) are still being
widely used in practice, since they are
considered conservative. This conception is
mainly based on the fact that the method
ignores the enhanced dissipation due to vertical
upward flow through the sand layer.
Nevertheless, the effect of the drain resistance
L may prove dangerous in the field, if the
improved layer is rather thick (Onoue 1988), or
if mixing of the drain material with the natural
soil occurs due to a vibratory construction
method (Boulanger et al 1998) and this even if
the selection of the drain material complies with
filter criteria.
It is well understood that gravel drains have
a composite beneficial operation in a sand layer
during an earthquake. As such, it has been
proven difficult to produce a design
methodology that takes into account all aspects
of the gravel drain operation. In this endeavor,
this paper presents the first results of employing
numerical analysis in the study of gravel drains,
a methodology that provides a means for
decoupling the various effects of their
composite operation. For this purpose, a
recently proposed bounding surface plasticity
model for non cohesive soil behavior under
cyclic loading is being used (Andrianopoulos et
al 2006) that has been implemented as a UserDefined-Model (UDM) routine in the 2D finite
difference code FLAC (Itasca Inc 1998). The
reliability of the numerical methodology is

verified via the simulation of a centrifuge test for


a uniform 19m-thick liquefiable sand layer that
was improved with gravel drains (Brennan and
Madabhushi 2006). Following this calibration,
this numerical methodology is used for
performing parametric analyses that attempt to
simulate the assumptions of Seed & Booker
(1977) and comment on their design
recommendations.
NUMERICAL METHODOLOGY: OUTLINE AND
VERIFICATION

The new constitutive model belongs to the


family of bounding surface plasticity models
with a vanished elastic region and incorporates
the framework of Critical State Theory. It is
based on a recently proposed model
(Papadimitriou et al., 2001; Papadimitriou &
Bouckovalas, 2002), which has been developed
with the goal to simulate the cyclic behavior of
non-cohesive soils (sands and silts), under any
(small-medium-large) cyclic shear strain
amplitude using a single (sand-specific) set of
constants, irrespective of the initial stress and
density conditions. The adoption of a vanished
elastic region differentiates the new model from
the original of Papadimitriou & Bouckovalas
(2002) and leads to a number of other
modifications as well, namely: (i) the
introduction of a new mapping rule and (ii)
modification of the existing interpolation rule.
However, the basic constitutive equations were
preserved, so that the reader can readily refer
to the original publications.
This new model was incorporated in the
code FLAC using its UDM capability. Integration
of the constitutive equations is performed with a
modified (two-step) Euler scheme. In order to
control the global integration error in the
computation of stresses, the sub-stepping
technique with automatic error control by Sloan
et al (2001) was adopted. Details on the
implementation can be found in Andrianopoulos
(2006) and Andrianopoulos et al (2006). The
capabilities of the new model have been
verified via extensive comparison with
laboratory element test results on Nevada sand
at relative densities of Dr = 40 & 60% and initial
effective stresses between 40 and 160 kPa
(Arulmoli et al, 1992). In particular, the
laboratory data originate from resonant column
tests as well as direct simple shear and triaxial
tests. It has been shown that accurate

468

3.0m. The natural sand material was Silica sand


(fraction E) at a relative density Dr 37% (e =
0.864), while the drain material consisted of
Silica sand (fraction B) that is characterized by
particle sizes 5 times larger, on average. The
centrifuge model was built inside a laminar box,
which was tested at 50g centrifugal acceleration
on the Cambridge 10m diameter beam
centrifuge. The base strong motion lasted
25sec and had a peak acceleration of 0.163g
and a predominant frequency of 0.82Hz (see
Figure 2). The response of the improved and
natural ground was monitored with the use of
six (6) accelerometers and sixteen (16) pore
pressure transducers.
The available material properties for both
fractions of Silica sand are presented in Table
1. Especially for fraction E, in lack of further
data, the constants of the constitutive model to
be used in the simulation were borrowed from a
pertinent calibration for Toyoura sand. The
reason for this selection is also explained in
Table 1, which shows that the grain size
distribution characteristics and the friction angle
of Silica sand fraction E resemble those of
Toyoura sand.

simulations with a single (sand-specific) set of


constants has been achieved and many basic
aspects of cyclic soil behavior, such as the
generation of excess pore pressures towards
liquefaction, permanent deformations, shearinduced dilation, softening and the effects of
evolving fabric anisotropy are well simulated
(Andrianopoulos, 2006).
The use of the new numerical methodology
for
liquefaction-related
boundary
value
problems has been extensively validated on the
basis of results from related centrifuge tests
(e.g. Andrianopoulos et al 2006, 2007).
Nevertheless, before using it for studying the
liquefaction mitigation capabilities of gravel
drains, it was considered necessary to validate
its performance for a pertinent centrifuge
experiment, namely the model test AJB-8a of
Brennan (2004).
The
test
arrangement
and
the
instrumentation for model test AJB-8a is shown
in Figure 1. In prototype scale, the experiment
refers to an almost 19m deep uniform sand
layer and the group of gravel drains reaches
down to the bottom of the layer. The group
consists of 17 drains of diameter d = 1.2m in a
triangular grid of a center to center distance s =

8915
8932

2259

6793
6794

6685

2976
8895
6797

6803

6669
6670

6679
9889
6677

6266

6788

16.8m

19m

1926

9882

6673

6783
8076

9m

6675

14m

16.8m

6.3m

Fig 1: Test arrangement of centrifuge test AJB-8a (Brennan 2004); Accelerometers are denoted by
quadrilaterals, whereas pore pressure transducers by dots

469

spectral acceleration (m/s2)

Recorded Motion
Filtered Motion

acceleration (m/s2)

-1

-2

0
0

10

time (s)

20

30

structural period T(s)

Fig 2: Recorded and filtered base motion during centrifuge test AJB-8a

Table 1: Material Properties of Silica sand (Fractions B & E, Brennan 2004) and Toyoura sand
Property

Silica sand fraction B


(drain material)

Silica sand fraction E


(natural sand)

D10 (mm)

0.095

D50 (mm)

0.140

0.15

D90 (mm)

0.150

emin

0.613

0.597

Toyoura sand

emax

1.014

0.977

Dr(%)

70.5

37.4%

37.4%

0.592

0.864

0.835

k (m/s)
o

cr ( )

-3

-4

7x10

10

32

31.2

the natural sand.


The grid used for the numerical analysis is
presented in Figure 4 and is 33.6m wide and
19m high. It consists of 589 elements (zones
in FLAC terminology) and is denser in the area
of the improvement. The five rows of cylindrical
drains of diameter d = 1.2m at a center-tocenter distance s = 3.0m are simulated in the
grid by five equivalent diaphragm walls of
thickness d = 0.6m, that have the same section
modulus with the actual rows of cylindrical
drains. This equivalence provides for the same
seismic ground response between the actual
3D geometry of improved ground and its
simulation with 2D plane strain analysis
(Papadimitriou et al 2006).

As an example of the calibration, Figure 3


presents the model prediction of a liquefaction
resistance curve for relative density Dr = 40%
and its comparison to respective data from the
literature. The drain material was simulated with
the use of a Ramberg-Osgood type non-linear
perfectly-hysteretic model, which has also been
implemented as a UDM routine in FLAC
(Andrianopoulos 2006). The reason for this
selection is that the drain material has a nonlinear behavior under strong seismic shaking,
but does not develop excess pore pressures, it
merely aids in the dissipation of those
developed in the neighboring sand. Compared
to the natural sand, the adopted calibration for
the drain material led to 50% larger initial elastic
moduli, a difference that widens during shaking
due to the reduction of the effective stresses of

470

CSR = c / vo

0.3

Nevada (Dr = 40%)


DeAlba et al (1976)
Hosono & Yoshimine (2004)
MODEL

0.2

0.1

0
1

10

number of cycles NL

100

Fig 3: Comparison of liquefaction resistance curve from model simulation for Dr = 37.4% (e = 0.864) with
related curves from the literature

Fig 4: Mesh used in the simulation of centrifuge test AJB-8a (shaded elements represent the gravel drains)

-10m, -19m) of the sand layer, within but also


outside the improved zone. Both recordings and
simulations show initial liquefaction (ru = 1.0) in
the upper 10m, whereas in deeper locations
significant excess pore pressures develop, but
not initial liquefaction (PPT 6673). The
improvement
does
not
prevent
initial
liquefaction, it only delays its appearance (at
t=7s in PPT2976, versus t=4s in PPT 6788) and
enables faster dissipation of excess pore
pressures after the end of shaking (not shown
in Figure 5). Overall the model simulations are
satisfactory, with the possible exception of the
somewhat smaller ru values at large depths,
and especially within the improved zone.
Figure 6 compares the time histories of
ground acceleration from the centrifuge test
recordings to their numerically simulated
counterparts, at various depths (-5m, -10m, 19m) of the sand layer and the agreement is
found satisfactory.

The horizontal acceleration time-history was


applied at the bottom of the mesh, after filtering
out the high frequency components of the
recorded motion (see Figure 2). The lateral
boundaries were tied to one-another in order to
ensure the same horizontal and vertical
displacements of the two boundaries, as
imposed by the laminar box device in the
centrifuge test. The vertical acceleration during
the experiments was minimal and was not taken
into account during the numerical simulation.
The permeability of the natural sand ks and the
drain material kd were given the values
measured in the laboratory (see Table 1), since
the pore fluid used was silicon oil that had a
viscocity 50cs to represent water in the
prototype scale.
Figure 5 compares the time histories of the
excess pore pressure ratio ru = u/vo from the
centrifuge test recordings to their numerically
simulated counterparts, at various depths (-5m,

471

1.2

0.8
0.4

ru

ru

1.2

PPT 6788

0.8
0.4

0
0

10

20

t(s)

30

Data
SIMULATION

10

t(s)

20

30

1.2

0.8

ru

ru

1.2
0.4

PPT 6266

0
0

10

20

t(s)

0.8
0.4

30

1.2

1.2

0.8

0.8

0.4

PPT 6673

0
0

10

20

t(s)

PPT 6679

ru

ru

PPT 2976

10

t(s)

20

0.4

30

PPT 6685

0
0

30

10

t(s)

16.8m

6788

2976

6266

6679

6685

20

30

19m

14m
9m

6673

2
1
0
-1
-2

t(s)

20

2
1
0
-1
-2

30

10

2
1
0
-1
-2

t(s)

20

10

t(s)

20

ACC 8915

10

20

30

ACC 8932

30

t(s)

2
1
0
-1
-2

30

ACC 8076

2
1
0
-1
-2

Data
SIMULATION

ACC 9889

a(m/s2)

10

a(m/s2)

a(m/s2)

a(m/s2)

ACC 8895

a(m/s2)

a(m/s2)

Fig 5: Comparison of data to simulations for the time history of the excess pore pressure ratio ru = u/vo
developed at various depths and locations of the AJB-8a centrifuge test

10

t(s)

2
1
0
-1
-2

20

30

ACC 9882

10

t(s)

16.8m

8895

8915

9889

8932

8076

9882

20

30

19m

14m
9m

Fig 6: Comparison of data to simulations for the time history of ground accelerations occurring at various
depths and locations of the AJB-8a centrifuge test

472

It is especially noteworthy that both the


recording and the simulation depict a
liquefaction-induced deamplification of the
acceleration at the ground surface, especially
outside the improved zone. In general, the
accelerations are more intense within the
improved zone, at all depths, and this has been
reproduced by the simulation. Yet, the
differences in the intensity of acceleration within
and outside the improved zone are not
significant and this because the group of drains
proved unable to prevent initial liquefaction, as
depicted by the ru time histories in Figure 5.

be purely horizontal towards the drains (Figure


7b), and vertical through the drains towards the
free surface.
If the intensity and the duration of the
seismic motion are sufficient, the encased sand
layer could reach liquefaction (ru = 1) after NL
loading cycles. According to Seed & Booker
(1977), after an improvement with gravel drains
of diameter a at an inter-distance b, the same,
horizontally drained, sand layer would reach a
maximum value of ru = ru,max < 1. As shown in
the design charts of Figure 8, this value of ru,max
is set to be a function of:
the improvement ratio, a/b,
the total number Ne of loading cycles of the
same intensity, normalized against NL, and
the dimensionless time factor Tad that
introduces the effects of the permeability of
the sand ks, its compressibility mv and the
duration of the seismic shaking td
Observe in Figures 7a and 7b that the
improved ground is simulated by using the
same mesh as that for the natural ground, but
by implementing different material properties for
elements of the drain material. In practical
terms, the response of the natural sand is
simulated using the new constitutive model and
the material constants used for Silica sand
fraction E above, whereas the response of the
drain material is simulated using the RambergOsgood type model and the material constants
used for Silica sand fraction B above. The clay
layers practically comprise an impermeable
boundary and should not affect the response of
the sand in any other manner. Therefore, they
are assumed to be a stiff elastic material (with
G = 300MPa).

PARAMETRIC NUMERICAL ANALYSES

Having established the reliability of the new


methodology, it is now possible to use it for the
study of the composite beneficial operation of
gravel drains in a sand layer during an
earthquake. In this effort, the first step is to
study the most commonly used design
methodology, i.e. that of Seed & Booker (1977).
For this purpose, the basic assumption of the
purely horizontal dissipation of excess pore
pressures is simulated with the use of the finite
difference grids shown in Figure 7. In particular,
a 1m thick sand layer is set to be encased
between 1m thick clay layers that have
permeability 1000 smaller than that of the sand
(this permeability difference was found sufficient
to disallow any vertical flow through the clay,
within the time frame of a seismic shaking).
Such an encased sand layer would develop
excess pore pressures under practically
undrained conditions during an earthquake
(Figure 7a), and if improved with gravel drains,
the dissipation of excess pore pressures would
65 kPa

1m
1m
1m
1m

65 kPa
1m
1m
1m
1m

CLAY

GRAVEL DRAIN

SAND

Fig 7: Mesh discretization and boundary conditions for reproducing the Seed & Booker (1977) assumptions

473

Tad=

4 ks td
mv a2 w

0.6
2
5

0.4

10
25
50
100

0.2

N / NL= 2

0.8

ru,max = umax/'

0.8

ru,max = umax/'

N / NL= 1

0.6

2
10

0.4
50

25

100

0.2
Tad=200

Tad=200
0

0
0

0.1

0.2

a/b

0.3

0.4

0.5

0.1

0.2

a/b

0.3

0.4

0.5

N / NL= 4

N / NL= 3
0.8

0.8
2

ru,max = umax/'

ru,max = umax/'

0.6
10
25

0.4
50
100

0.2

2
5

0.6
10
0.4

25
50
100

0.2

Tad=200

Tad=200
0

0
0

0.1

0.2

a/b

0.3

0.4

0.5

0.1

0.2

a/b

0.3

0.4

0.5

Fig 8: Design charts of Seed & Booker (1977) for groups of gravel drains against liquefaction

initial effective stresses. Finally, to ensure


continuous saturation of the soil deposit during
shaking, the ground water table was set to be
1m higher than the ground surface.
The seismic motion was applied at the base
of the mesh and it was assumed to be a
harmonic sinusoidal motion with a peak
acceleration of 4m/s2 at a period T=0.3s.
Besides simplicity, the reason for applying a
sinusoidal motion and not an actual seismic
motion recording in our analyses was the need
to duplicate another basic assumption of the
methodology of Seed & Booker (1977), i.e. the
fact that they based their empirical curve for the
rate of excess pore pressure buildup on
element laboratory tests, which apply
successive cycles with a constant cyclic stress
ratio CSR under undrained conditions. As such,
the analysis for fully undrained conditions (with
the mesh of Figure 7a) provided a practically
uniform response throughout the sand layer,
whose rate of excess pore pressure build up is
presented in Figure 9.

In terms of permeability, the sand is


assumed to have a ks = 10-4m/s (as that of
Silica sand fraction E), the clay kc = 10-7m/s
(ks/kc = 1000), whereas the drain material was
assigned a reference value of kd = 2x10-2m/s
(kd/ks = 200).
Furthermore, note that the grid was
intentionally chosen to be low-rise (3m high
versus 31m wide) in order to avoid amplification
of the horizontal motion and/or potential rocking
(i.e. generation of vertical vibration for a purely
horizontal base motion). The lateral boundaries
were tied to one-another in order to ensure the
same horizontal and vertical displacements of
the two boundaries, as imposed by a laminar
box device in a centrifuge test. This type of
lateral boundary conditions have proven
successful in simulating the actual vibration of
natural deposits and this is why they were
chosen for these analyses. To ensure a
numerically stable response of the very surficial
sand layer (1 2m deep) during shaking, an
additional surcharge of 65kPa was applied at
the ground surface in order to increase the

474

1
Undrained Response
range within mesh
0.8

0.6

ru

Seed et al. (1976)


Upper limit

0.4
Seed et al. (1976)
Average

0.2

0
0

0.2

0.4

N / NL

0.6

0.8

Fig 8: Rate of excess pore pressure buildup from the analysis for fully undrained conditions and comparison
with the empirical findings of Seed et al (1976)
1

no drainage
(a/b = 0)

0.8

0.6

ru

a/b = 0.2

initial
liquefaction

0.4
a/b = 0.3
0.2
a/b = 0.5
0
0

0.2

0.4

0.6

N / NL

0.8

1.2

Fig 9: Rate of excess pore pressure buildup from analyses for various improvement ratios a/b and
comparison with the analysis for fully undrained conditions (a/b = 0, of Fig. 8)

group of gravel drains of diameter d = 0.8m set


at a spacing s = 2.7m, according to
Papadimitriou et al (2006). Hence, the
improvement
geometry
of
Figure
7b
corresponds roughly to an improvement ratio
a/b = 0.3. In addition, two extra analyses were
performed for less and more dense grids of
gravel drains, i.e. with a/b = 0.2 and 0.5,
respectively. The numerically predicted time
histories of the excess pore pressure ratio ru for
all improvement ratios (a/b = 0, 0.2, 0.3 and
0.5) are presented in Figure 10. As observed,

The comparison of Figure 8 shows that the


constitutive model simulations for fully
undrained conditions are in accordance with the
empirical findings of Seed et al (1976) that
served as basis for the design charts of Seed &
Booker (1977).
In the sequel, the analysis with the mesh of
Figure 7b was performed. The thickness d of
the equivalent gravel drain diaphragm walls in
Figure 7b is equal to 0.33m and they are at a
center-to-center distance of 2.7m. This 2D
improvement geometry corresponds to a 3D

475

4NL. The comparison presented in Figure 11


shows that the numerical predictions are quite
consistent with the analytical ones for N/NL = 1
(and various a/b values). However, for N/NL = 2,
3 and 4 the difference between numerical and
analytical predictions increases, since the
former are not affected (since N/NL is no longer
a problem parameter, as explained above),
while the latter move towards higher ru,max
values, in general.
Attempting to explain this very important
difference the following may be stated:
For design purposes, the important factor is
the maximum ru value reached during
shaking (ru,max). Hence, Seed & Booker
(1977) defined ru,max as the ever-current
maximum value of ru from the beginning of
shaking. As such, an ru,max value can either
increase or remain constant with time, but
cannot ever decrease. A careful examination
of the analytical predictions for N/NL=4
shows that these are lower than those for
N/NL = 3, a fact that is physically impossible.
More quantitative examples of this
discrepancy can be found in Papadimitriou et
al (2007).
The Seed & Booker (1977) charts imply an
increasing ru value with time, at least up until
N/NL = 3, whereas the numerical analyses
show an ru value reaching its peak sometime
before N/NL = 1. This difference can be
attributed to assumption (c) of the Seed &
Booker (1977) work, i.e. the fact that the
drains are set to dissipate a rate of ru buildup
from the fully undrained tests, while this rate
was never reached in the improved ground
due to the drainage initiating from the start of
the shaking.

the curve for a/b = 0 (undrained conditions)


holds as the maximum possible rate of excess
pore pressure buildup. As the improvement
ratio a/b increases, the excess pore pressure
ratios decrease due to drainage. More
importantly for this study, the maximum value of
ru = ru,max decreases as the improvement ratio
a/b increases, and this is in qualitative
accordance with the design charts of Seed &
Booker (1977) shown in Figure 8.
However, there is also an important
differentiation presented in this figure. Namely,
a detailed study of the results of Figure 10
shows that the maximum value of ru in all
analyses is attained sometime within the 0 <
N/NL 1 time window, after which the ru slowly
reduces, on average, with time, even though
the seismic shaking continues. This qualitative
result has been shown to hold in all performed
analyses, the presentation of which is beyond
the scope of this paper, and implies that even if
the seismic shaking had many more cycles than
NL (i.e. N/NL > 1) this would not affect the ru,max
value at all. In other words, this figure shows
that the N/NL ratio is not a problem parameter,
as long as N/NL > 1.0.
This indirect conclusion is verified in Figure
11, which compares the ru,max variation with a/b,
obtained from the numerical analyses and the
Seed & Booker (1977) analytical solution. More
specifically, the triangles in Figure 11 denote
the ru,max values of the numerical analyses
presented in Figure 10. The circles in Figure 11
correspond to analytical estimates (on the basis
of the Seed & Booker (1977) solution), which
correspond to the site-earthquake conditions
assumed in the numerical analyses. These
conditions correspond to the following four (4)
distinct site-earthquake combinations:
N/NL = 1
&
Tad = 5.7
N/NL = 2
&
Tad = 11.4
N/NL = 3
&
Tad = 17.1
N/NL = 4
&
Tad = 22.8
Observe that if Tad = 5.7 for an earthquake with
N = NL and the shaking continues for up to N =
2NL, then its corresponding value of Tad also
doubles (Tad = 11.4), because it should
correspond to an earthquake of a double
duration. In this respect, the foregoing four (4)
site-earthquake combinations are interrelated
and this is why the respective numerical
analyses were performed with a single
simulation run per a/b, that lasted up to N =

CONCLUSIONS

This paper presents a well-calibrated


numerical methodology for evaluating the
effectiveness of a group of gravel drains for
mitigating earthquake-induced liquefaction. In
an in-depth study of the available design
methods, the first step is the evaluation of the
cornerstone of the relative literature, i.e. the
design charts of Seed & Booker (1977).

476

N / NL= 1

0.8

ru,max = umax/'

0.8

ru,max = umax/'

N / NL= 2

analytical
(Tad = 5.7)

0.6

0.4

0.2

analytical
(Tad = 11.4)

0.6

0.4

0.2

numerical

numerical

0
0

0.1

0.2

a/b

0.3

0.4

0.5

0.1

0.2

a/b

0.3

0.5

N / NL= 4

N / NL= 3
0.8

0.6

analytical
(Tad = 17.1)

0.4

0.2

ru,max = umax/'

0.8

ru,max = umax/'

0.4

0.6

analytical
(Tad = 22.8)

0.4

0.2

numerical

numerical

0
0

0.1

0.2

a/b

0.3

0.4

0.5

0.1

0.2

a/b

0.3

0.4

0.5

Fig 11: Design charts of Seed & Booker (1977) for groups of gravel drains against liquefaction

By
performing
parametric
numerical
analyses in an effort to duplicate the
assumptions of their work, it is shown that the
general form of their analytical solution is
verified, as well as its quantitative accuracy for
relatively short duration events with N = NL.
However, for relatively longer events (with N >
NL) it appears that the analytical solution is
over-conservative
and
may
lead
to
unnecessarily dense groups of gravel drains.
One approximate way to avoid this is to design
the grid of gravel drains based on the analytical
solution for N/NL = 1.0, even in cases where the
number of significant (uniform) cycles of
shaking exceeds NL.
Further study is obviously needed (and is
currently underway), before the foregoing
general conclusions for the methodology of
Seed & Booker (1977) can be quantified and
used in practice. It is important to clarify here
that the aforementioned over-conservatism of
the Seed & Booker (1977) charts applies only to
cases with an insignificant drain resistance L,
i.e. cases where the drain permeability kd is

much larger than that of the natural sand ks and


the thickness of the sand layer is small. On the
other hand, when the L is significant, the Seed
& Booker (1977) charts are usually not used
and alternative design methods are preferred
(e.g. Onoue 1998). Since these methods inherit
assumption (c) of the Seed & Booker (1977)
method, they also inherit the depicted overconservatism that is assumption induces.
Hence, the foregoing general conclusions may
be considered applicable to all design methods
that are based on the foregoing assumption.
ACKNOWLEDGEMENTS

The work presented herein has been


financially supported by the General Secretariat
for Research and Technology (....) of
Greece, through research project
23 (X-SOILS). This contribution is gratefully
acknowledged. The authors would also like to
thank Dr. Andrew Brennan from the University
of Dundee UK for providing the electronic data
from his centrifuge test.

477

Iai, S. and Koizumi, K. (1986) Estimation of


earthquake induced excess pore pressure for
th
Japan
gravel drains, Proceedings, 7
Earthquake Symposium, 679-684
Itasca Consulting Group, Inc. (1998) FLAC Fast
Lagrangian Analysis of Continua, Version 3.4,
Users Manual, Minneapolis: Itasca
Matsubara, K., Mihara, M. and Tsujita M. (1988)
Analysis of gravel drain against liquefaction
and its application to design, Proceedings, 9th
World Conference on Earthquake Engineering,
Vol. III 249-254
Papadimitriou, A.G., Bouckovalas, G.D. and Dafalias,
Y.F. (2001) Plasticity model for sand under
small and large cyclic strains, Journal of
Geotechnical
and
Geoenvironmental
Engineering, ASCE, 127, no. 11, 973-983
Papadimitriou, A.G. and Bouckovalas, G.D. (2002)
Plasticity model for sand under small and
large cyclic strains: a multiaxial formulation,
Soil Dynamics and Earthquake Engineering,
22, 191-204
Papadimitriou,
A.G.,
Moutsopoulou,
M.-E,
Bouckovalas, G.D., and Brennan, A.J. (2007)
Numerical
investigation
of
liquefaction
mitigation using gravel drains, Proceedings,
th
4 International Conference on Earthquake
Geotechnical Engineering, paper 1548,
Thessaloniki, Greece, June 25 22
Papadimitriou,
A.G.,
Vytiniotis,
A.C.
and
Bouckovalas, G.D. (2006) Equivalence
between 2D and 3D numerical analyses of the
seismic response of improved sites,
Proceedings, 8th US National Conference on
Earthquake Engineering, paper 2020, San
Francisco, CA, April 18 22
Seed, H.B., and Booker, J.R. (1977) Stabilization of
potentially liquefiable sand deposits using
gravel drains, Journal of the Geotechnical
Engineering Division, ASCE, 103(GT7): 757768
Seed, H.B., Martin, P.P., and Lysmer, J. (1976)
Pore water pressure changes during soil
liquefaction, Journal of the Geotechnical
Engineering Division, ASCE, 102(GT4): 323346
Sloan, S.W., Abbo A. J., and Sheng D. (2001)
Refined explicit integration of elastoplastic
models with automatic error control,
Engineering Computations, 18 (1/2), 121-154
Onoue, A. (1988) Diagrams considering well
resistance for designing spacing ratio of gravel
drains, Soils and Foundations, 28(3): 160-168

REFERENCES
Andrianopoulos, K.I. (2006) Numerical modeling of
static and dynamic behavior of elastoplastic
soils, Doctorate Thesis, Department of
Geotechnical Engineering, School of Civil
Engineering, National Technical University of
Athens (in Greek)
Andrianopoulos, K.I., Papadimitriou, A.G., and
Bouckovalas, G.D. (2006) Implementation of a
bounding surface model for seismic response
th
of sands, Proceedings, 4 International FLAC
Symposium on Numerical Modeling in
Geomechanics, May
Andrianopoulos, K.I., Papadimitriou, A.G., and
Bouckovalas, G.D. (2007) Use of a new
bounding surface model for the analysis of
earthquake-induced liquefaction phenomena,
Proceedings, 4th International Conference on
Earthquake Geotechnical Engineering, June
Arulmoli, K., Muraleetharan, K.K., Hossain, M.M.
and Fruth, L.S. (1992) VELACS: verification of
liquefaction analyses by centrifuge studies;
Laboratory Testing Program Soil Data
Report, Research Report, The Earth
Technology Corporation
Boulanger, R.W., Idriss, I.M., Stewart, D.P.,
Hashash, Y., and Schmidt, B. (1998)
Drainage capacity of stone columns or gravel
drains for mitigating liquefaction, Geotechnical
Special Publication No.75, Proceedings,
Specialty
Conference
on
Geotechnical
Earthquake Engineering and Soil Dynamics III,
Seattle Washington, ASCE, Vol.1, 678-690
Brennan, A.J. (2004) Liquefaction remediation by
vertical drain groups, PhD Thesis, University
of Cambridge UK
Brennan, A.J. and Madabhushi, S.P.G. (2006)
Liquefaction remediation by vertical drains
with varying penetration depths, Soil
Dynamics and Earthquake Engineering,
26(5):469-475
DeAlba, P., Seed, H.B. and Chan, C.K. (1976) Sand
liquefaction in large-scale simple shear tests,
Journal of the Geotechnical Engineering
Division, ASCE, 102(GT9): 909-927
Han, J. and Ye, S.-L. (2001) Simplified method for
consolidation rate of stone columns reinforced
foundations, Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, 127(7):
597-603
Hosono, Y. and Yoshimine, M. (2004) Liquefaction
of sand in simple shear condition,
Proceedings, International Conference on
Cyclic Behaviour of Soils and Liquefaction
Phenomena, 129-136, Bochum, Germany,
March 31 April 2

478

Estimating earthquake induced settlements on granular soils


Application to shallow foundations
D. N. Egglezos1,
1

National Technical University of Athens, Free Lancer, Consultant of Greek Ministry of Culture, Greece

Abstract
The article presents a set of general relationships for permanent volumetric strain from cyclic
loading. These relations can be used to predict earthquake induced settlements on granular soils,
with direct application to shallow foundations. Volumetric strain is calculated at the end of each
loading cycle for the following cases: drained or undrained conditions, isotropic initial stress state
(1=3, e.g. free field) or anisotropic initial stress state (13, e.g. initial static shear), constant
cyclic stress or constant cyclic shear. The variables of the general relations are based on statistical
analysis of experimental data from drained cyclic triaxial tests and relate directly to the initial stress
or strain state and density of the soil element. The accuracy of predictions is examined through
comparison with a) experimental data b) well established charts from other researchers and c)
measured settlements due to earthquake, from reported cases histories.

More specifically, these relations apply to


dry or saturated sands, constant stress or
constant strain cyclic loading, free field stress
state (that is absence of static shear) or stress
state with initial shear.
The proposed relations result from drained
cyclic triaxial tests (21 tests on fine
Oosterschelde sand) and for this reason they
apply initially to triaxial stress state. However,
the extension to more general stress states
through adequate assumptions is easily
attained.
The proposed empirical relations provide the
means
for
simplified
computations
of
earthquake induced settlements involved in
design of bearing capacity of foundations, thrust
on retaining structures and so on. In addition,
they can be used as reference for calibration of
constitutive models aimed at the prediction of
cyclic response of granular soil.

INTRODUCTION

Densification of granular soils subjected to


seismic loading is a well known phenomenon
responsible for settlements of structures. This
densification in case of dry soils takes place
simultaneously with dynamic loading. In case of
saturated soils the development of excess pore
pressure from undrained cyclic loading does not
allow initially the occurence of settlements.
Indeed, settlements occur at post earthquake
phase during the dissipation of induced excess
pore pressures. The rate of these settlements
depend on permeability characteristics of soil
involved.
Although the importance of settlement
calculation for the safe design of structures is
obvious the usual practice is the omit of any
cosideration, mainly due to lack of simple
means for this. Indeed, for a simplified design,
at present, the available means are limited to
anyway useful- charts from reported literature
(Ishihara,1992,etc. ).
The scope of this work is the proposal of
simple empirical relations for calculation of
volumetric strains of sand from cyclic loading,
applying to usual initial stress state of soil
involved in geotechnical design.

VOLUMETRIC STRAIN MODEL FOR SANDS


Data base for volumetric strain model on sands

The empirical relations for volumetric strain on


sand resulted from adequate statistical fitting
upon sufficient experimental data from drained
cyclic
triaxial
tests
on
uniform
fine

479

qc=1dc/2,

Oosterschelde sand with grain size d50=0.17mm


and uniformity coefficient Cu=1.40 (Lambe,
1979).
These tests are consolidated anisotropically
in compression (CAD tests, 1>3). Cyclic
stress (double) amplitude 1dc remained
constant during testing.
The range of the initial stress and density
parameters are presented in Table 1, as
function of the following invariant magnitudes:
p=(1+2 3)/3,
qo=(1-3)/2,

P=qo/(p ), CSR=1dc/(2p) or

c(%) (double amplitude cyclic strain)


where, M denotes the slope of PTL line (e.g.
Ishihara et al., (1975), Luong and Sidaner
(1981)), which separates contractive from
dilative behaviour in space qo-p.

Table 1: Range of initial parameters in drained cyclic triaxial tests on Oosterschelde sand
Test
number

CSR

c(%)

p/pa

Initial
void ratio e

d50 (mm)

Uniformity
coefficient Cu

21

0.06 0.25

0.020-0.124

1.43-4.81

0.30-0.77

0.70-0.77

0.17

1.40

Volumetric strain on dry sands (drained


loading):
Figure 1 shows in log-log scale the typical
development of volumetric strain of sand when
it is sheared cyclically in state of free
drainage(continuous line) with number of
loading cycles (data come from a typical CAD
triaxial test with constant cyclic stress).
This behaviour can successfully be simulated
from the following formula (dashed line):

VOL (N) = VOL 1 N C VOL ,max

relating
to
maximum
relative
density
(DR(%)=100).
Analytical
prediction
(just
qualitative approach for better comparison with
data) based on equation 1 is also included In
Figure 1 with dashed line.
The volumetric strain at the end of 1st loading
cycle VOL1 in equation 1 (either for constant
stress or for constant strain) can be expressed
(as it has been shown in previous research of
the author (Egglezos (2004)) as simple product
of exponential terms:

(1)

for constant stress:

Indeed the development of volumetric strain


in the log-log scale is practically linear (that is
exponential). In equation 1, VOL is the
volumetric strain (%), N is the number of cycles,
c express the effect of loading cycles and VOL,
max the nominal upper limit of volumetric strain:

VOL , max =

eo emin
1 + eo

a2

VOL 1

p'
= ACSR o e a 3 f (P ) VOL , max
pa
a1

where, f (P ) =

(2)

(3a)

1 P
a4
1 (0.8P )

while, for constant strain:

1.00

experimental data (Oosterschelde sand)

VOL 1 = B C e b 3
b1

VOL (%)

Typical prediction (eq. 1)

1 P
VOL , max
b4
1 (0.8P )

(3b)

0.10

where, pa is the atmospheric pressure for


normalisation of units (pa =100 kPa).
It is worth noting that the above equations
3a and 3b are not arbitrary formulations of VOL1
but result from adequate simplification of strict
analytical equations (Egglezos, 2004). These
analytical formulae are based on Rowes
constitutive equation for volumetric strain

TEST D196
c = 0.39
0.01
1

10

100

1000

NUMBER OF CYCLES,

Fig 1: Typical development of volumetric strain with


number of cycles

480

statistical analysis of data.


Application of equations 4a and 4b to
prediction of exponent c from experimental data
leads to the following range of prediction ratio
Rc (=cpredicted/cmeasured): 0.85<Rc<1.15 either for
constant stress or constant strain. These values
justify the adoption of equations 4a and 4b for
c-exponent calculation.

(Rowe, 1962) and the hyperbolic model for -


description (e.g. Duncan, (1970)). In addition,
the variables of these formulae result directly
from the initial stress and density state of soil.
Constants in equations 3a and 3b result
from
multivariable
non-linear
regression
statistical analyses (StatSoft Inc., 1995) upon
experimental data of Table 1. Indeed, the
statistical analyses leads to a very good fitting
of VOL1 formulae on experimental data. The
values of these constants are presented in
Table 2.

Development of volumetric strain on saturated


sands (undrained loading)

The extension of empirical relations in cases


with restrained drainage (undrained loading)
can be done according to Rowes equation
(Rowe, 1962) for description of volumetric
strain

Table 2: Constants of the empirical relations


Constant stress CSR: No of data=21, R=0.921
A
a1
a2
a3*
0.770
1.55
0.744
5.700
Constant shear c(%): No of data=20, R=0.946
B
b1
b2**
b3*
2.372
1.17, c<1%
0.0
6.470
0.163, c>1%

a4*
0.578

dvol = dpo/t + dP = dpo/t +dvol

b4*
0.432

where,
vol=total volumetric strain, po=mean
effective stress (1+23)/3, P=permanent
shear strain, vol= permanent volumetric strain,
t=tangential Bulk Modulus
In case with restrained drainage during
cyclic loading where dvol=0, equation 5,
transforms to:

* Values of constants a3, b3 and a4, b4 are based on the relevant


effect on excess pore pressure accumulation
** The effect of mean effective stress is negligible (b2=0.035) and
for this reason it is omitted in equation 3a.

Exponent c, (in eq. 1) which express the


effect of number of loading cycles, in the case
of excess pore pressure build up, can be
expressed statistically (123 data, R=0.60) from
initial state parameters as simple product of
exponential terms (Egglezos (2004):

dvol = (d-dU)/t+dvol=0 dU=t dvol

(4a)

for constant strain:

c( c (% )) = 0.833e1.295 C0.060 exp(0.480 P)

(6)

From the above equation results direct


relation of excess pore pressure and volumetric
strain in undrained loading. This (restrained
during cyclic loading) volumetric strain may
produce post earthquake settlement
when
dissipation of excess pore pressure takes
place.
For this reason it seems rational that
formulae initially developed for prediction of
excess pore pressure from cyclic loading
(Egglezos, 2004, Egglezos (2007)) can offer a
general pattern for prediction of postearthquake volumetric strain.

for constant stress:

c(CSR ) = 1.07e1.58 CSR 0.202 exp(0.521P)

(5)

(4b)

Although these formulae result from


statistical analyses of undrained cyclic triaxial
tests on sand, and describe the effect of cycles
to excess pore pressure build up, it is assumed
that can be used equivalently for the effect of
loading
cycles
on
volumetric
strain
accumulation. This is in agreement to Rowes
constitutive equation (eq.5) for linear relation of
excess pore pressure and volumetric strain (eq.
6) as it is explained in next paragraph. It is
worth noting that the limited range of c
exponent from the available data from drained
cyclic tests does not allow for a reliable

a. CIU conditions (qo=0) constant stress:

For CIU conditions and constant stress


(which simulate well free field conditions) the
excess pore pressure of sand after N cycles of
cyclic loading can be expressed (according to
previous research of the author (Egglezos,
2004, Egglezos, 2007)) as follows :
b

U
U(N) = U 1 N C 1 + 0.01 N

max
N 1st

481

(7)

EVALUATION OF EMPIRICAL RELATIONS

In equation 7, U is the excess pore


pressure, N is the number of cycles, N1st
corresponds to the number of cycles for the end
of 1st stage of excess pore pressure
accumulation (N1st relates to excess pore
pressure U1st), c express the effect of loading
cycles at the 1st stage of excess pore pressure
accumulation (for N<N1st), b express the effect
of loading cycles at the final (2nd) stage, U1 is
the excess pore pressure at the end of 1st cycle
and Umax
the upper limit in excess pore
pressure directly estimated from initial state
parameters: U max = p ' o q o M
For CIU tests Umax= po and corresponds to
stress state (q, p) on PTL line for loading
cycles required for initial liquefaction (N=NL).
U1st and N1st in granular soils are expressed
with high accuracy as constant ratios of Umax
and NL respectively (Egglezos (2004), Egglezos
(2007)): U1st = 0.52 Umax, 1st = 0.54 L, while
the exponent b can be taken with mean value
5.80 (typical range of b exponent from exact
calculation: 4.80<b<6.80, Egglezos (2004)).
According to equation 7, volumetric strain in
the case examined seems rational that could be
expressed as:

N
VOL (N) = VOL 1 N 1 + 0.01

N 1st
C

The evaluation of the proposed empirical


relations for volumetric strain of granular soils is
based on the following applications:
Prediction of experimental data from drained
cyclic tests

Figure 2 shows the application of empirical


relations for calculation of volumetric strain in
sands, from cyclic loading with constant shear
strain of great amplitude (c>1%). The
experimenal data are coming from 4 drained
cyclic torsional tests on Toyoura sand (Hyodo et
al. (2002)).
The basic initial parameters of the data are
included in Table 3. In figure 2 experimental
data are presented with continuous line while
predictions (with use of equations 1, 3b and 4b)
with dashed line. It is observed the relatively
good fitting of predictions on experimental data.
More specifically the prediction ratio RVOL
ranges for number of cycles N=1-10 from 0.56
to 1.48 (0.56<RVOL<1.48):
where, R VOL =

VOL , max (8)

VOL , predicted
VOL ,measured

In addition, figure 3 shows the application of


empirical relations for calculation of volumetric
strain in sands, from cyclic loading with
constant shear stress. The experimental data
are coming from 4 drained cyclic torsional tests
on Ham River sand (Tsomokos, 2005). The
basic initial parameters of experimental data are
also included in Table 3. It is observed the
relatively good fitting of predictions (from
equations 1, 3a and 4a) on two cases (3a, 3d).
In the other cases it is observed underprediction for the initial cycles, while accuracy is
improved with increasing number of cycles. The
required CSRTX for application of empirical
relations refers to triaxial loading. For this
reason reported CSR values (which refer to
torsional shear) are transformed as follows:
CSRTOR = CSRTX cr
where, cr is adequate correction factor for the
stress state (e.g. Kastro (1975)).
To be noted that experimental data used for
evaluation of the empirical relations have not
been used for the statistical evaluation of
constants in equations 3a and 3b (type A
prediction).

The above formula for post cyclic volumetric


strain is modified adequately -comparing to that
for drained loading- to account for the sharp
increase of pore pressure after the first stage of
excess pore pressure accumulation.
b. CIU conditions (qo=0) constant strain, CAU
conditions (qo>0) - constant stress or strain:

In the above cases excess pore pressure is


expressed (Egglezos (2004)) as:

U(N) = U 1 N C U max
This formulation for pore pressure is
identical to that for volumetric strain in drained
conditions. For this reason equation 1:

VOL (N) = VOL1 N C VOL ,max


is considered adequate for prediction of
volumetric strain in the cases examined in this
paragraph.

482

10.00

100
TEST 161

(a)

(b)

VOL(%)

VOL(%)

TEST 164

1.00

10

Equation 1
EASURED

0.10

10

NUMBER OF CYCLES, N

100

100

100

(d)

(c)

TEST 172

VOL(%)

VOL(%)

10

NUMBER OF CYCLES,

100
TEST 165

10

10

NUMBER OF CYCLES,

10

100

10

100

NUMBER OF CYCLES,

Figure 2: Prediction of volumetric strains of cyclic torsional tests in Toyoura sand


1.0

Tsomokos, 2005

1.00

test in HRS

VOL(%)

VOL(%)

Equation 1

0.1

0.10

Tsomokos, 2005

test in HRS

(b)

(a)
0.0

0.01
1

10
100
1000
NUMBER OF CYCLES

10.0

1.00
Tsomokos, 2005

(c)

test in HRS

VOL(%)

VOL(%)

10
100
1000
NUMBER OF CYCLES

1.0
Tsomokos, 2005

0.10

(d)

test in HRS

0.1

0.01
1

10
100
1000
NUMBER OF CYCLES N

10
100
NUMBER OF CYCLES N

Figure 3: Prediction of volumetric strains of cyclic torsional tests in Ham River sand

483

1000

Table 3: Initial parameters in cyclic CID torsional tests (evaluation of the empirical relations)
Drained cyclic torsional tests in Toyoura sand (constant shear)
Test
p(kPa)
c(%)
eo
161
98
6
0.76
164
98
6
0.89
165
70
6
0.89
172
134
6
0.89
Drained cyclic torsional tests in Ham River sand (constant stress)
Test
p(kPa)
CSR
eo
HRS-1
130
0.144
0.715
HRS-2
130
0.292
0.704
HRS-3
130
0.371
0.708
HRS-4
130
0.128
0.702

P
-

PI

P
-

PI

The continuous lines correspond to reported


values of required CSR for a M=7.5 earthquake
(for development of given volumetric values) as
function of SPT blows (Tokimatsu & Seed,
(1987)).
Prediction of required CSR value is obtained
from empirical equations 8 and 3a:

Comparison with literature curves for volumetric


strain of sands

i) Figure 4 shows application of the


proposed empirical relations for prediction of
cyclic stress CSR required for development of
predetermined values of volumetric strain:
VOL= 0.1%, 0.2%, 0.5% and 1%
respectively.
0.50

0.50
TOKIMATSU & SEED, 1987 (a)

0.40

Equation 9

0.30

0.30

VOL=0.1(%)

0.20

0.20

0.10
0.00

(b)

VOL=0.2(%)

CSR

CSR

0.40

0.10
0

10

20

0.00

30

10

N1,60

20

30

N 1,60

0.60

0.60
(d)

(c)
VOL=1.0(%)

VOL=0.5(%)

CSR

0.40

CSR

0.40

0.20

0.00

0.20

10

N1,60

20

0.00

30

10

N1,60

20

Figure 4: Comparison of empirical relations with Tokimatsu and Seed curves (1987)

484

30

CSRTX

0.774

N
p'
= VOL (N )/ 0.77eeq5.70 o N c 1 + 0.01

N 1st
a

5.80

(11.55 )

N160
5
10
15
20
22
25
28
30
32
35

(9)
The application is based in the following
assumptions:
1. Void ratio required for the application (eeq)
is obtained as function of relative density DReq
for assumed margin values emax=1.0 and
emin=0.5 which consist typical void ratio limits for
uniform clean sand: eeq=1-0.5*Dreq
2. The relative density Dreq has been
obtained as function of SPT blows 160
(owles,1995):

Dreq
0.10
0.30
0.46
0.60
0.63
0.68
0.74
0.78
0.81
0.85

3. For magnitude M=7.5 earthquake the


number of equivalent uniform cycles is
eq(M)=16

1E+1

1E+1
(a)

1E+0

Dr=45%

VOL(%)

VOL (%)

1E+0

(b)

1E-1
Tokimatsu & Seed (1987)

1E-2

1E-2

1E-1

1E+0

1E-1
1E-2

Equation 1,

1E-3
1E-3

Dr=60%

1E-3
1E-3

1E+1

1E-2

c(%)

1E+0
1E-1
VOL(%)

1E-1

1E+0

1E+1

c(%)
(c)
Dr=80%

1E-2
1E-3
1E-4
1E-3

1E-2

1E-1

1E+0

1E+1

c(%)

Figure 5: Comparison of empirical relations with Tokimatsu and Seed curves (1987)

relevant value for free field condition CRRff is


obtained from the following transformation:

4. Initial effective vertical stress: po=100


kPa.
5. the required value for c-exponent in
equation 3a is obtained iteratively through
equations 4a and 9.
To be noted that the calculated CSRTX from
equation 9 refers to triaxial loading. The

CRRff = 0.90 CRR cr,


where cr is adequate correction factor for the
stress state (e.g. Kastro (1975)). Practically, for
the specific application: CSRTX=2.0 CSRff.

485

From inspection of chart it is observed a


satisfactory fitting of predictions (dashed lines)
with reported values of CSRff (continuous lines)
particularly for volumetric strain ranging 0.10.5%.

The continuous lines correspond to reported


values of Tokimatsu and Seed curves (1987)
for volumetric strain of dry sands while dashed
lines refer to proposed empirical relations
(equations 1, 3b and 4b). The void ratio
required for empirical relations is obtained as it
was explained in case I of this paragraph.
From inspection of figure 5 it is observed a
satisfactory fitting of predictions with Tokimatsu
curves (continuous lines) with prediction ratio
ranging 0.40<RVOL<2.0 for 0.01%c1%.

ii) Figure 5 shows application of the


proposed empirical relations for prediction of
volumetric strain from cyclic loading of constant
shear strain at different levels of relative
densities (Dr=45%, 60% 80%), for an
M=7.5 earthquake.

2.00
M=6 (Neq=6), M=7 (Neq=12), M=7.5 (Neq=20)

1.80

Equation 7&8, DR(%)=50, (Neq=6, 12, 20)


Ishihara & Yoshimine (1992), DR(%)=50
Equation 7&8, DR(%)=90, (Neq=6, 12, 20)

FSL

1.60

Ishihara & Yoshimine (1992), DR(%)=90

1.40
1.20
1.00
Neq =6

Neq =20

Neq =12

Neq =6

Neq =12

Neq =20

0.80
0.00

0.20

0.40
VOL(%)

0.60

0.80

Figure 6: Application of empirical relations for prediction of volumetric strain in saturated sands
comparison with Ishihara & Yoshimine chart (1992)

iii) Comparison with Ishihara and Yoshimine


chart for calculation of volumetric strain of
saturated sands for constant stress cyclic
loading. Figure 6 shows application of the
empirical relations for prediction of volumetric
strain as function of safety factor against
liquefaction FSL. The continuous curves
correspond to volumetric strain from application
of the empirical relations for densities DR=50%
and
DR=90%
and
earthquakes
with
magnitudes, M=6 (Neq=6), M=7 (Neq=12) and
M=8 (Neq=20). The void ratio is obtained as it
was explained in case I of this paragraph. In the
same figure are included, with dashed line,
Ishihara & Yoshimine curves for the relevant

densities. Generally there is good agreement of


the two kind of curves.
However it is worth noting that empirical
relations are more detailed as they take
account not only for soil density (as it happens
to Ishihara & Yoshimine charts) but in addition
for the earthquake magnitude (through the
effect of number of equivalent cycles).
Prediction of settlements comparison with
measured settlements in case histories

This paragraph comprises comparison


between predictions from the proposed
empirical relations and measured settlements.

486

i) Settlement prediction of saturated sand


in Marina district of San Francisco from
Loma Prieta earthquake (Kramer, 1995).
Prediction is based on equations 1, 3b and 4b.
In addition, multidirectional effect of the shaking

has been taken in account by doubling the


calculated VOL. Analytically, the required steps
for calculation of settlements are included in
Table 4.

Table 4: Settlement calculation of saturated loose sand fill in Marina district of San Francisco
3
(Loma Prieta earthquake (1989): M=6 /4, Neq=10)
Layer

thickness
(kPa)
(m)
1
55
1.8
67
1.5
82
1.7
96

160

Dreq

eeq

CSRff

CSRtx

VOL,tx

VOL,,max

(m)

7.5
11.4
8.6
18.7

0.357
0.426
0.356
0.508

0.821
0.786
0.821
0.745

0.141
0.165
0.180
0.186

0.282
0.330
0.360
0.372

0.606
0.585
0.637
0.550

0.001815
0.002243
0.080883
0.001508

0.176471
0.160526
0.176588
0.140661

0.001815
0.004037
0.121325
0.002564
0.13
0.12

PREDICTED SETTLEMENT OF SATURATED LOOSE SAND FILL (m)


MEASURED SETTLEMENT (m)

ii) settlement of dry sand deposit from


San Fernando earthquake (M=6.6, Neq=9,
Kramer (1995)). Prediction is based on
equations 8, 3a and 4a. The required steps for
calculation of settlements are analogous to the
previous case (see Table 4). From application
of empirical relations a settlement of 8cm is
estimated. This value compares well to
measured settlements ranging from 7 to 10 cm.

performed for, the center of the model building,


the edge of the model building, and for the free
field far enough from the model building. The
induced (effective) vertical and shear stresses
on the subsoil are estimated through analytical
equations from elastic theory (Steinbrenner,
1934). The nominal dynamic excitation at the
base of experimental arrangement was a 0.25g
sinusoidal pulse with ten uniform cycles. For
calculation of CSR in different depths, the
average contact pressure from the Centrifuge
model building, was obtained equal to 150 kPa.
The required geotechnical information for
prediction is obtained from relevant literature
(Arulmoli K. et al., 1992). Fig. 7 shows the
comparison between measured settlement of
model building during the test and predictions,
for the loading cycles of test (N=1-10).
Continuous black line refers to 10 uniform
cycles of constant acceleration (ah=0.25g) while
the dashed line refers to increased acceleration
(ah=0.32g) for cycles 1 to 4 and ah=0.25g for
cycles 4-10. In the last case (which is closer to
the recorded acceleration within soil mass
below the foundation of the model building)
volumetric strain at the ith cycle (i4) is
calculated as follows:
VOL(N=i) = VOL(N=4) + VOL(N=i, ah=0.25g)
- VOL(N=4, ah=0.25g)

iii) Prediction of settlement of the backfill


in Lefkada Marina from the 14-8-2003
earthquake, M=6.4, Neq=8 (Alexoudi et al.
(2006), Anastasiadis et al.(2006), Gazetas et al.
(2006)). The required geotechnical information
for prediction is obtained from borehole section
in the Marina area reported in literature
(Anastasiadis,(2006)). The required steps for
calculation of settlements are analogous to
case i (see Table 4). From application of
empirical relations a settlement of 17cm is
estimated. This value compares well to
measured settlements in various sites of
Marina, ranging from 10 to 20 cm.
iv) Prediction of shallow foundation
settlement recorded during the VELACS Model
Test No 12 (Arulmoli K. et al., 1992). The test
was performed in centrifuge device and
simulates the response of sandy subsoil
(Nevada sand of DR(%)=60) under undrained
cyclic loading, a) in the vicinity of a centrifuge
model building with shallow foundation, and b)
in the free field. The calculations of settlements
with the proposed empirical relations are

where, VOL(N=4) = max

{[

VOL(N=4,ah=0.32g)+VOL(N=4,ah=0.25g)

],

[VOL(N=4,ah=0.32g)+VOL(N=3,ah=0.32g)]

487

8.4 cm, while at the edge of the model building


ranges from 0.67 cm to 0.78 cm, for the two
loading cases examined respectively.

For this last case, accumulation of


settlement in time fits particularly well to
recorded settlement.
It is also worthmentioning that predicted
settlement at free field ranges from 7.0 cm to

Settlement: m

-0.05
0.00
0.05
0.10

Measured
Equation 8, ah=0.25g, Neq=1-10

0.15

Equation 8, ah=0.32g (N=1-4) & ah=0.25g (N=4-10)

0.20
0

4
6
8
Number of cycles: N

10

12

Figure 7: Predicted and measured settlements from VELACS 12 Model Centrifuge Test

CONCLUSIONS

e) The parameters of the empirical relations


result directly from the initial stress (or strain)
and density state of soil under examination.
f) The values of constants in the empirical
relations result from statistical analysis of
available experimental data for drained cyclic
triaxial tests (21 tests on Oosterschelde sand).
g) The extension of empirical relations to
undrained conditions is achieved through the
general constitutive relation between excess
pore pressure and volumetric strain (Rowe,
1962).
h) The accuracy of the proposed relations is
tested with application on a) prediction of
impartial experimental data from
cyclic
torsional tests on sand (type A prediction), b)
comparison with well established charts from
relevant literature for calculation of volumetric
strains and c) calculation of post-earthquake
settlements measured in a number of case
histories. In addition, for application in
settlement
calculation
of
shallow
foundations, due to seismic loading,
predictions from empirical relations compare to
recorded settlement from centrifuge model test.

A The major conclusions of this research


are the following:
a) For the present time no other empirical
relation for simplified calculation of volumetric
strain in granular soils is referred on reported
literature.
b) Geotechnical designers, as a usual
practice, omit calculation of volumetric strain in
granular soils because of lack of simplified
methods and absence of relevant provision in
many national anti- seismic codes.
c) The proposed empirical relations for
prediction of volumetric strain on sands with
adequate transformation (concept of equivalent
void ratio: Egglezos (2001), Egglezos(2007))
may also apply in case of gravelly or silty soils.
d) The proposed model predicts the
accumulated volumetric strain at the end of
each cycle of cyclic loading. It is characterized
of great mathematical simplicity and permits the
estimation of volumetric strain for different initial
stress or strain states of soil (constant cyclic
stress or shear, initial static shear or not,
drained or undrained conditions e.t.c.).

488

In all the above cases predictions agree


particularly well with measurements.
Finally as main conclusion it has to be
mentioned that the proposed empirical relations
for volumetric strain can be used for simplified
calculation of earthquake settlements in
geotechnical design, as well as in calibration of
numerical codes for prediction of dynamic
behavior of granular soils. Of course, the
extension of experimental data base in future,
with additional data from cyclic tests, could
further improve the accuracy of predictions.

Soils in Liquefaction Hazard Evaluation, Proc.


th
5 Hhellenic Conference on Geotechnical and
Geoenviromental Engineering, Xanthi, Greece,
(in Greek)
Egglezos D.N. (2007) "Empirical relations for
earthquake pore pressure build-up in gravel,
th
Proc. 4 ICEGE, Thessaloniki, Greece, (it has
been accepted for presentation)
Gazetas G. et al. (2006) Failure of harbor quaywalls
th
in the Lefkada 14-8-2003 earthquake, Proc. 5
Hellenic Conference on Geotechnical and
Geoenviromental Engineering, Xanthi, Greece,
(in Greek)
Ishihara K.F., Tatsuoka and Yashuda S. (1992)
Undrained deformation and liquefaction of sand
under cyclic stresses, Soils and Foundations,
32(1), 173-188

REFERENCES
Arulmoli K., et al.(1992) VELACS: verification of
liquefaction analyses by centrifuge studies;
Laboratory Testing Program Soil Data Report,
Research Report, The Earth Technology
Corporation

Ishihara K.F., and Yoshimine M. (1975) Undrained


deformation and liquefaction of sand under
cyclic stresses, Soils and Foundations, 16(1),
1-16
Kramer S.L. (1996) Geotechnical earthquake
engineering, Prentice Hall International Series in
Civil Engineering and Engineering Mechanics,
Upper Sddle River, New Jersey.

Alexoudi M. et al.(2006) The influence of site-effects


in the seismic assessment of water systems.
th
The case of Lefkas,
Proc. 5 Hellenic
Conference
on
Geotechnical
and
Geoenviromental Engineering, Xanthi, Greece,
(in Greek)
Anastasiadis A.I. et al. (2006) The Lefkas
earthquake (M=6.2, Aug. 14, 2003). Strong
ground motion and valuation of subsoils impact,
th
Proc. 5 Hellenic Conference on Geotechnical
and
Geoenviromental Engineering, Xanthi,
Greece, (in Greek).
De Alba P. et al. (1976) Sand liquefaction in large
scale simple shear tests, Journal of the
Geotechnical Engineering Division, ASCE
102(9), 155-163

Lambe T.W. (1979) Cyclic triaxial tests on


Oosterschelde sand, MIT Research Report
R79-24, Soils Publication No. 646
Luong M. P. and Sidaner J.F. (1981) Comportment
th
cyclique et transitoire des sables, Proc. 10
ICSMFE, Stockholm, Sweden, 3, 257-260
Norwegian Geotechnical Institute (1988) Bearing
capacity of gravity platform foundations on
sand, Report 52422-5
O

Duncan J. M., Chang C. Y. (1970) Nonlinear


analysis of stress and strain in soils, Journal of
Geotechnical Engineering Division, ASCE, 96(5),
pp. 1629-1653
Egglezos D.N. (2001) "Prediction of excess pore
pressures on silty sands and sandy silts from
th
cyclic loading, Proc. 4 Hhellenic Conference
on Geotechnical Engineering, Athens, Greece,
(in Greek)

Rourke T.D. et al. (1991) Lifeline and


geotechnical aspects of the 1989 Loma Prieta
nd
erthquake, Proc. 2
International Conference
on
Recent
Advances
in
Geotechnical
Earthquake Engineering and Soil Dynamics, St.
Louis, Missouri, Vol. 2, 1601-1612

Pierce, W. G. (1983) Constitutive relation of


saturared sand under undrained loading, Ph. D.
Dissertation, Dept. of Civil Engineering,
Rensselaer Polytechnic Institute, Troy, New
York
Pyke R., Seed H.B. and Chan C.K. (1975)
Settlements of sands under multi-directional
loading,
Journal
of
the
Geotechnical
Engineering Division, ASCE, Vol. 101, GT4,
379-398

Egglezos D.N. (2004) Experimental and theoretical


investigation of soil behavior under cyclic
loading, Thesis submitted in partial satisfaction
of the requirements for the degree of PhD in
Engineering, N.T.U.A., Greece, (in Greek)

Rowe P.W. (1962) The stress-dilatancy relation for


static equilibrium of an assembly of particles in

Egglezos D.N. (2006) "Application of Empirical


Relations for Excess Pore Pressure of Granular

489

contact, Proc. Royal Society, Vol. A269, 500527,.


Seed H.B. and Silver M.L. (1972) Settlements of dry
sands during earthquakes, Journal of the Soil
Mechanics and Foundation Division, ASCE,
98(SM4), 381-397
Seed H.B. & De Alba P. (1986) Use of SPT and
CPT tests for evaluating the liquefaction
resistance of soils. Proc., Insitu Testing 86,
ASCE
Silver M.L and Seed H.B. (1971) Volume changes
in sands during cyclic loading, Journal of the
Soil Mechanics and Foundation Division, ASCE,
97(SM9), 1171-1188
StatSoft, Inc. (1995) STATISTICA for windows,
Computer Program, Tulsa, USA
Steinbrenner, W. (1934) Tafeln zur Setzungberechnung, Die Strasse, Vol. 1, p. 121
Sangseom, J. (1988) The behavior of silt under
triaxial loading, Thesis submitted in partial
satisfaction of the requirements for the degree of
Master of Science in Engineering, Davis,
California, USA
The Earth Technology Corporation (1992) VELACS
(Verification of analyses by centrifuge studies),
Laboratory testing program, Soil data report,
Earth Technology Project No. 90-0562
Tokimatsu K. and Seed H.B. (1987) Evaluation of
settlement in sand due to earthquake shaking,
Journal of Geotechnical Engineering, ASCE
113(8), 861-878
Tsomokos A.I. (2005) Experimental study of the
behaviour of a soil element under monotonic
and cyclic torsional shear, Ph. D. Dissertation,
Dept. of Civil Engineering, N.T.U.A., Athens,
Greece, (in greek)

490

Unstable Behaviour of a Fine Sand Under Cyclic Loading


V. N. Georgiannou, A. Tsomokos
National Technical University of Athens, Greece

Abstract
The behaviour of a fine sand under cyclic torsional loading is investigated in the hollow
cylinder apparatus. During cyclic loading unstable response is initiated along the instability
line defined under monotonic loading. Unstable response is associated with a sudden
increase in the rate of shear strain and excess pore water pressure accumulation. The
influence of the number of cycles on the development of excess pore pressure up to the
initiation of unstable response can be described by empirical equations. The liquefaction
characteristics of the sand are presented.
INTRODUCTION

which consists of fine and rounded particles


with mean particle size D50 of 0.22mm. The
sand has emin=0.540, emax=0.865 (determined
following
the
method
described
by
Kolbuszewski (1948)) and Gs=2.64. Its grain
size distribution curve is shown in Figure 1. All
loose specimens were formed by pluviation
through water following the procedure
suggested by Bishop and Henkel (1957) for the
preparation of sand. After confirming saturation,
with B values in excess of 0.97, specimens
were isotropically consolidated to a range of
effective stresses, pi=(1+2+3)/3. An ageing
period of 12 hours preceded shearing. The tests
were performed in the hollow cylinder apparatus
of the National Technical University of Athens.

The stress-strain behaviour of sands under


cyclic loading has been the subject of various
studies in the past. Cyclic strength of sands has
been generally evaluated by means of
undrained cyclic triaxial tests in which the
principal stress directions are interchanged.
This cyclic stress condition is not appropriate to
simulate the idealized field stress conditions
during earthquake loading which are more
accurately represented by torsional shear tests.
Symes et al. (1984) and Shibuya et al.
(2003) cyclically rotated the principal stresses
acting on hollow cylinder specimens of a loose
fine sand between =0 and 90 from their
initial vertical and horizontal directions. They
observed that partial liquefaction was initiated
along the effective stress path defined under
monotonic loading at constant . In torsional
cyclic shear tests performed on isotropically
consolidated hollow cylinder specimens, the
major principal stress is rotated by 45 from
the vertical on reversing the direction of the
applied shear stress.
In this paper the onset of unstable response
under cyclic loading conditions is investigated
for a fine sand (Fontainebleau sand). Instability
marks a drastic change in the response of a
sand since it is accompanied by a sudden loss
of strength and the development of large pore
water pressures and shear strains.

Percentage Finer (%)

100
80
60
40
20
0
0.01

0.1
1
Particle Size (mm)

10

Fig 1: Grain size distribution curve of Fontainebleau


sand.

MATERIAL AND TEST METHODS

A general view of the apparatus is given in


Figure 2. Torsional shearing was applied in this
study under the same internal and external

Fontainebleau sand is a uniform silica sand

491

pressure (pi=po) resulting in b=sin2, where the


parameter b=(2-3)/(1-3) expresses the
influence of the intermediate principal stress
and is the rotation of the major principal stress
direction from the vertical. The axial load W was
close to zero during the tests. The specimens
had an outer diameter of 70mm, an inner
diameter of 40mm and a height of
approximately 140mm.

and is the double amplitude rotational angle


of the specimen, in rad.
CYCLIC
SAND

BEHAVIOUR

OF

FONTAINEBLEAU

The undrained response of Fontainebleau


sand to cyclic loading was investigated for loose
specimens having void ratios in the range
0.713-0.743 as indicated in Table 1. Denser
specimens have also been subjected to cyclic
loading. After isotropic consolidation a cyclic
shear stress amplitude SA is applied to the
specimens corresponding to a range of cyclic
stress ratios CSR=0.44-0.97, where CSR is the
ratio of the cyclic shear stress to the undrained
shear strength (at transient peak) measured in a
monotonic torsional shear test at the same
consolidation pressure.
Table 1. Specimen characteristics

Fig 2: Hollow cylinder apparatus.

Tests were controlled and interpreted in


terms of average stresses and strains given by
the following expressions after Hight et

al.(1983),

Z =

3M T
=
2 (r03 ri3 )

and

SA

Loading
Type

ei

SA/ pi

CSR

NIL

F2

*M

0.731

F10

0.713

0.2387

0.97

F11

0.714

0.2079

0.86

F12

0.736

0.1816

0.74

F13

0.732

0.1612

0.66

24

F14

0.721

0.1440

0.58

48

F15

0.743

0.1211

0.50

68

F16

0.734

0.1050

0.44

150

0.665- 0.1392- 0.460.700 0.2200 0.71


*M, C = monotonic, cyclic loading

19513

F20F24

2(r03 ri3 )
, where MT is the applied
3H(r02 ri2 )

Typical results representing all cyclic


torsional hollow cylinder tests performed on
loose specimens at different cyclic stress ratios
are presented next. Figure 3 illustrates the
effective stress paths followed by two
specimens of Fontainebleau sand at similar void
ratio isotropically consolidated to the same
mean effective stress pi=130kPa and
subsequently subjected to monotonic (F2) and
cyclic (F13, CSR=0.66) torsional loading,
respectively. Under monotonic loading the
shear stress, z, steadily increases to a
transient peak value which marks the initiation
of unstable response, since with further loading

torque, r0 and ri are the current outer and inner


radii, the circumferential angular displacement
and H the initial height of the specimen. Cyclic
loading was applied at a frequency of 0.1Hz.
During cyclic loading the single amplitude shear
stress
and
strain
were
defined
as

TR + TL
and
2 (r + ri2 )(r0 ri )
(r0 + ri )
=
, where TR, TL are the single
4H

SA =

Test

2
0

amplitude torque in the clockwise and


anticlockwise direction, both positive in value

492

80

'=38o

a)

PTL

40

z=2.2%

'PTL=33o

z=0.5%

'IL=22
2

Shear stress z(kPa)

IL
o

0
0

40

80

120

160

8
6

3
7

-40

Mean effective stress p' (kPa)

Shear strain z (%)

15

b)

10
5

0
-5

-10

2
3

4
50

100

150

Time (sec)

200

250

-15

Excess pore water pressure u(kPa)

-80

140

c)

8
6

120
100

80

60
40

20
2

0
0

50

100
150
Time (sec)

200

250

Fig 3: Undrained monotonic (F2) and cyclic (F13) torsional hollow cylinder test on Fontainebleau sand: a)
effective stress paths; b) shear strain against time curve-cyclic loading; c) excess pore water pressure
against time curve-cyclic loading.

halted by the effective stress path reaching the


point of phase transformation1. As soon as the
specimen reaches the phase transformation line
dilative tendencies take over and the effective
stress path follows the sands failure envelope.
The instability line (Lade (1993)) and the phase
transformation line determined from undrained
tests (IL=22, PTL=33) have also been
plotted in Figure 3(a). The instability line
connects the peak points of the effective stress
paths and the origin of the stress space. The

the shear strength of the specimens drops


rapidly to a minimum (Figure 3(a)) as the rate of
excess pore water pressure and shear strain
accumulation increases to a maximum value. It
is worth noting that initiation of instability occurs
at small shear strains (of the order of 0.5%)
Similar observations have been reported in the
literature by various researchers (e.g. Vaid and
Chern (1985), Alarcon-Guzman et al. (1988),
Ishihara (1993), Georgiannou (2006)). The
reduction in strength is followed by a quasisteady state stage (QSS, Ishihara (1993))
during which the specimen deforms under a
nearly constant shear stress. This phase is

Phase transformation is taken as the change from contractant to


dilatant tendencies (Ishihara et al. (1975))

493

pressure accumulation is arrested at the phase


transformation line defined under monotonic
loading. Changing the loading direction from
point 3 leads the specimen to initial liquefaction.

slope of the instability line seems to be uniquely


determined for specimens of similar void ratio
irrespective of their initial effective stress level
and/or the small variations in the value of void
ratio considered herein and is considerably
smaller compared to the slope of the phase
transformation line.
During the first quarter of the first loading
cycle (points 1-2 in Figure 3(a)) the cyclic stress
path follows initially but doesnt coincide with
the monotonic stress path. Unloading from point
2, which corresponds to maximum positive
shear stress, and further loading in the opposite
direction leads to the accumulation of excess
pore water pressure (point 3). Unstable
behaviour of the specimen is introduced when
the cyclic effective stress path reaches the
instability line defined from the monotonic tests
(point 4) during the first quarter of the 22nd
loading cycle. Loading from point 4 to 5 leads to
a sudden increase of excess pore water
pressure (Figure 3(c)) and shear strain (Figure
3(b)). At point 4 the shear strain value is only
0.5% but increases rapidly to approximately
2.2% at point 5. Unloading from point 5 and
reloading in the opposite direction to point 6
leads to further accumulation of excess pore
water pressure, while the shear strain value at
point 6 is close to zero. Point 6 is a phase
transformation point (PTP), for cyclic loading
conditions, where contractive are replaced by
dilative tendencies. The response of the
specimen suggests that the excess pore water
pressure becomes equal to the confining
pressure only once the stress ratio (point 7 in
Figure 3(a)) is greater than the value of stress
ratio at phase transformation defined under
monotonic loading and loading changes
direction (from point 7 to 8). At point 8 the
specimen looses its strength and liquefies.
However, it deliquefies with further loading and
its shear stress climbs up the failure envelope at
the expense of large strains (9%) showing
cyclic mobility.
Figure 4 includes a cyclic test (F11) loaded
at CSR=0.86 and it shows more clearly the
cyclic path after unstable response is introduced
when the effective stress path approaches the
instability line (point 1 in Fig. 4). High excess
pore pressure and shear strain are accumulated
for further loading (point 1 to 2) and the cyclic
stress path migrates towards the origin, while it
remains parallel to the effective stress path
under monotonic loading, until the excess pore

80

'=38o
PTL
'PTL=33o

40

IL

'IL=22o

Shear stress z (kPa)

2
1

0
0

40

80

120

160

-40

Mean effective stress p' (kPa)

-80

Fig 4: Undrained monotonic (F2) and cyclic (F11)


torsional hollow cylinder test on Fontainebleau sand.

Unstable response is associated with a


sudden increase in the rate of shear strain and
excess pore water pressure accumulation. The
influence of the number of cycles on the
development of excess pore pressure is shown
in Fig. 5(a) for five specimens of loose
Fontainebleau sand tested at CSR=0.44-0.74.
Each point on each curve represents the excess
pore water pressure at the end of a loading
cycle. Similar response was observed for other
fine sands (Georgiannou & Tsomokos, 2007).
The following equation (1), which is slightly
modified from the equation proposed by
Ishibashi et al. (1977), can fit the characteristic
shape of these curves up to the initiation of
unstable response as indicated by the curves
based on the empirical equation and shown in
Figure 5(a) as broken lines.

u(N)
p'i

= (1

u(N1) c1 c2 *N
SA(N) c5
) *(
)*(
) (1)
c3
p'i
p
'
(N1)
N +c
4

In equation (1) N is the current number of


loading cycles, u(N) and p(N) are the excess pore
water pressure and the mean effective stress at
the end of loading cycle N, u(N)= u(N)-u(N-1),
SA(N) is the cyclic shear stress during cycle
number N, pi is the initial mean effective stress

494

and c1, c2, c3, c4, c5 are constants depending on


the sand, its density and cyclic loading
conditions i.e. constant or varying cyclic shear
stress, SA. For the tests considered herein the
values of the constants are given in Table 2.

to bring the specimen to initial liquefaction NIL.


In Figure 5(b) all cyclic tests performed on
Fontainebleau sand, as described in Table 1,
have been included. The normalized results
form a narrow band for all cycles prior to initial
liquefaction.

Table 2. Values of constants


c1

c2

c3

c4

c5

loose

-0.5

3.80

1.66

-0.21

2.40

dense

-0.8

1.65

1.66

-0.55

2.40

0.40
0.35
0.30
(SA / pi') / f(ei)

Fontainebleau sand

In Figure 5(b) the excess pore pressure at


the end of each cycle has been normalized with
respect to initial mean effective stress (u/pi)
and plotted against the current cycle number
normalized with the number of cycles required

0.25
0.20
0.15

a = 0.218
b =-0.152

0.10
0.05

SA / pi'=a*f(ei)*N ILb

0.00
1

10

Correlation coefficient R=0.916

100

NIL

1000

Fig 6: Liquefaction curve for Fontainebleau sand.


1,0

a)

The number of cycles required to bring the


specimens of Fontainebleau sand to initial
liquefaction has been plotted in Figure 6 against
cyclic shear stress normalized with respect to

0,8

u/pi'

0,6

initial mean effective stress, (


0,4

of two densities (Dr=565% and 425%) were


consolidated to the same mean effective stress
(pi=130kPa) and subsequently subjected to a
range of CSR=(0.44 to 0.97). In Figure 6 the
results of the tests have been normalized with
respect
to
the
void
ratio

CSR=0.44, ei=0.734
CSR=0.50, ei=0.743

0,2

CSR=0.58, ei=0.721
CSR=0.66, ei=0.732
CSR=0.74, ei=0.736
Empirical equation

0,0
1

10

100

1000

Number of cycles N
1,0

function, f (e i ) =

Fontainebleau sand

0,8

SA
) . Specimens
p' i

(2.17 e i ) 2
, proposed by
(1 + e i )

Hardin & Richart (1963). The mean trend can


be described by the following mathematical
equation (2):

b)

0,6
u/pi'

SA
= a * f (e i ) * N bIL
p' i

0,4

where SA is the cyclic shear stress, pi is the


mean effective stress, NIL is the number of
cycles required to cause initial liquefaction, f(ei)
is the void ratio function and a, b are constants.

0,2

0,0
0,0

0,2

0,4

0,6

0,8

(2)

1,0

N/NIL

CONCLUSIONS

Figure 5: Excess pore water pressure accumulation


with number of cycles: (a) loose Fontainebleau sand; (b) normalized excess pore
water pressure curves for loose and dense
Fontainebleau sand.

1. Under
monotonic
torsional
loading
Fontainebleau sand exhibits loss of strength
after a transient peak value which lies on the
instability line (IL).

495

shapes and sizes, Geotechnique, Vol.56, No.9,


pp. 639-649.

2. Under undrained cyclic torsional loading


instability is initiated when the cyclic effective
stress path reaches the instability line
defined from the monotonic torsional hollow
cylinder tests at a shear strain of 0.5% or
less.
3. The development of excess pore pressure
during cyclic loading including the onset of
instability can be predicted by empirical
equations and can be normalized for all
cycles prior to initial liquefaction.
4. The liquefaction characteristics of loose and
dense
Fontainebleau
sand
can
be
normalized with respect to the void ratio
function.

Georgiannou, V.N. and Tsomokos A. (2007),


Comparison of two fine sands under torsional
loading,
Submitted
to
the
Canadian
Geotechnical Journal.
Ishibashi I., Sherif, M.A., and Tsuchiya, C. (1977),
Pore-pressure rise mechanism and soil
liquefaction, Soils and Foundations. Vol.17,
No.2, pp. 17-27.
Ishihara, K. (1993), Liquefaction and flow failure
during earthquakes, Geotechnique, Vol.43, No.
3, pp. 351-415.
Ishihara, K., Tatsuoka, F., and Yasuda, S. (1975),
Undrained deformation and liquefaction of sand
under cyclic stresses, Soils and Foundations,
Vol.15, No.1, pp. 29-44.

REFERENCES

Kolbuszewski, J.J. (1948), An experimental study of


the maximum and minimum porosities of sands,
2nd ICSMFE, Rotterdam, Vol.1, pp. 158-165.

Alarcon-Guzman, A., Leonards, G.A., and Chameau,


J.L. (1988), Undrained monotonic and cyclic
strength of sands, J. Geotech. Engrg, ASCE,
Vol.114, No.10, pp. 1089-1109.

Lade, P.V. (1993), Initiation of static instability in the


submarine
Nerlerk
berm,
Canadian
Geotechnical Journal, Vol.30, No.6, pp. 895-904.

Bishop, A.W., and Henkel, D.J. (1957), The


measurement of soil properties in the triaxial
test, London: Edward Arnold.

Shibuya, S., Hight, D.W., and Jardine, R.J. (2003),


Four-dimensional local boundary surface of an
isotropically consolidated loose sand, Soils and
Foundations, Vol.43, No.2, pp. 89-103.

Hardin, B.O. and Richart, F.E. (1963), Elastic wave


velocities in granular sands: measurements and
parameter effects, J. SMFD, ASCE, Vol.89,
No.1, pp. 33-65.

Symes, M.P.R., Gens, A., and Hight, D.W. (1984),


Undrained anisotropy and principal stress
rotation in saturated sand, Geotehnique, Vol.34,
No.1, pp. 1-27.

Hight, D.W., Gens, A., Symes, M.J. (1983), The


development of a new hollow cylinder apparatus
for investigating the effects of principal stress
rotation in soils, Geotechnique, Vol.33, No.4, pp.
355-383.

Vaid, Y.P., and Chern, J.C. (1985), Cyclic and


monotonic undrained response of saturated
sands, Advances in the art of testing soils under
cyclic conditions, ASCE Annual Convention,
Detroit, Michigan, pp. 120-147.

Georgiannou, V.N. (2006), The undrained response


of sands with additions of particles of various

496

Numerical Simulation of Liquefied Sand


Using a Shear-Thinning Fluid Model
Kay-Uwe Schenkengel, Andreas Becker, Christos Vrettos
Technical University of Kaiserslautern, Germany

Abstract
Lattice Boltzmann equations are becoming a very popular simulation tool in fluid dynamics.
Macroscopic quantities like fluid density, velocity or stress, are expressed as moments of a
distribution function. This approach may be manipulated to simulate non-Newtonian fluids
with shear strain-rate dependent viscosities. Laboratory tests on liquefied sand show a
shear-thinning behaviour, i.e. decreasing viscosity as shear strain rate increases. Similar
observations are made for muds and other geophysical fluids. A nonlinear Carreau-type
rheological model with upper and lower limiting values of the fluid viscosity is adopted as
constitutive relation, and the effects of solidification are included. The main features of the
numerical procedure developed are described in the paper. Typical application problems
such as the collapse of a liquefied slope, and the flow around a wall in liquefiable soil are
then solved enabling the qualitative assessment of the mechanisms involved.

INTRODUCTION

a numerical method that considers the phase


transformation in a simplified manner. A fluid
dynamics approach has been implemented by
Uzuoka et al. (1998).
The phase transition from solid to liquid as
well as the realistic description of the two
phases in a truly coupled procedure is still very
difficult. A practical alternative at this stage of
investigation is to decouple the two states.
Among the available numerical procedures the
lattice Boltzmann method has become during
the last years a standard method for treating
engineering problems in different fields of
computational fluid dynamics, Succi (2001). In
particular, the flow of non-Newtonian fluids like
blood (Artoli & Sequeira, 2006) or melted metal
(Krner at al., 2003) have recently been solved
by applying this method.
In this paper we will take the first step
towards the application of the lattice Boltzmann
method to simulate the behavior of liquefied
sand. We assume that the liquefied sand is a
non-Newtonian fluid with nonlinear viscosity.
The rheological model adopted here is a shearthinning fluid of the Carreau type.

The onset of liquefaction has been the


subject of extensive investigations during the
past decades. In the meantime a number of
empirical methods are available to designers to
predict the occurrence of this effect. However,
there are several cases where liquefaction can
only be prevented by costly countermeasures
or even has to be considered in the design of
structures, such as pile foundations. Thus,
during the last years research is focusing in
estimating the behavior of liquefied soil.
Some empirical relations have been
proposed, e.g. prediction of the undrained
shear strength from SPT blow count numbers,
Seed & Harder (1990). Parallel to this work, a
number of constitutive models have been
proposed to describe this medium and the
transformation from a solid to a fluid.
Still there is no consensus among scientists
on the proper description of the material. A
number of experimental results indicate that a
viscous fluid is an appropriate practicable
model, Kawakami et al. (1994), de Alba &
Ballestero (2006).
In a recent work Sato et al. (2002) described

497

THE LATTICE BOLTZMANN METHOD

For the D2Q9 model the velocity vectors


with velocity c are

Historically the lattice Boltzmann method is


a successor of a special kind of cellular
automata, the so-called lattice gas automata,
Wolf-Gladrow (2000). It could, however, be
regarded as a discretization of the Boltzmann
equation and therefore an approximation of the
Navier-Stokes equation. In He & Luo (1997) this
has been shown on the basis of the BGK model
(Bhatnagar, Gross and Krook, 1954) that is a
simplified version of the Boltzmann equation:
wf & wf
[ &
wt
wx

1
 (f  f ( 0 ) )
O

&

^ei `

1
(0)
 (fi  fi ), i
O

(1)

0,...,N

(3)

where the lattice constant velocity


c 'x / 't is given in dependence of the grid
size 'x , the time step 't .
&
Note, that e0 c(0,0)T represents particles
at rest while the other particles move to their
neighbor positions within the grid.
With further discretization of time and space
via a finite difference scheme we obtain the
lattice Boltzmann equation:

Equation (1) describes the evolution


in time
&&
t of a density distribution function f ( x, [,t ) in the
phase space. The left side of equation (1) is the
advective transport term with the velocity vector
&
&
[ , and the position vector x . The right hand
side represents the relaxation towards the
Maxwell-Boltzmann equilibrium distribution
function f (0) with the relaxation time O .
To solve for f numerically, equation (1) is
first discretized in the velocity
& space using a
finite set of velocity vectors ei (i 0,...,N ) what
leads to the velocity discrete Boltzmann
equation,

wfi & wfi


 ei &
wt
wx

0 1 0 1 0 1 1 1 1

c
0 0 1 0 1 1 1 1 1

& &
fi ( x  ei 't , t  't )

&
&
1
eq &
fi ( x, t )  (fi ( x, t )  fi ( x, t ))
W

(4)

By rearranging terms of this fundamental


form we can see the two main steps of the
lattice Boltzmann procedure. The streaming
step, in which the distribution functions fi are
moved to the adjacent grid cells, and the
collision step, in which the relaxation towards
the local equilibrium takes place:
& &
&
fi ( x  ei 't ,t  't )  fi ( x, t )


strea min g

(2)

&
1 &
 (fi ( x,t )  fieq ( x,t ))
W

(5)

collision

In the above equation W

O / 't is the

dimensionless relaxation time, and fi eq is the


discrete equilibrium distribution function. The
latter is an approximation of the Maxwell(0 )
Boltzmann distribution function fi valid for low
Mach numbers:

&
In equation (2) fi ( x,t ) is equivalent to
& &
f ( x,ei ,t ) . The discretization of the velocity
space leads to a Cartesian grid. Fig 1 shows
the form and the numbering of local directions
and represents a so-called D2Q9 model, the
latter denoting D = 2 dimensions and Q = 9
velocities according to the naming scheme
proposed by Qian et al. (1992).

fi

eq

3 & &
9 & &

w i U 1  2 ei u  4 (ei u )2
c
2c

3 & &

u u
2c 2

(6)

&
where u is the macroscopic velocity of the
fluid particles, U is the fluid density. The
weighting factors w i are obtained by a
Chapmann-Enskog analysis (Wolf-Gladrow,
2000):

Fig 1: The D2Q9 grid

498

4 / 9 ,

1 / 9 ,
1 / 36 ,

i
i
i

0
1, 2 , 3 , 4
5, 6, 7,8

the local computability of the stress and strain


tensor. There is no need to estimate any
velocity gradients. The momentum flux tensor
can directly be obtained from the nonequilibrium part of the distribution functions.
This allows the use of a wide range of
rheological models. As mentioned in the
Introduction the Carreau-Yasuda model is
selected here for the analysis.

(7)

The mass and momentum density are given


in terms of moments of distribution functions
calculated in the discretized velocity space:
8

i 0 fi i 0 fi eq
&
8 &
8 & eq
U u i 0 ei fi i 0 ei fi

(8)

(9)

( W  1/ 2) cs2 't

(10)

with
c/ 3

(11)

1000
Viscosity coefficient [Pa*s]

cs

denoting the speed of sound in the fluid. The


pressure in the fluid medium is then given by
p U cs2 .
&
Gravitational force g (0, g ) is introduced
&
via an increase of macroscopic velocity u :

&
'u

&
Wg

(13)

where Q f , and Q0 are the limiting values of


viscosity at low and high values of shear strain
rate J , and N , a , and b are material
parameters.
The suitability of this equation to describe
real behavior of liquefied sand is demonstrated
in Fig 2 through comparison with experimental
results taken from Kawakami et al. (1994).

The kinematic viscosity of the fluid medium


Q is a result of the Chapmann-Enskog
expansion mentioned above:

Q f  ( Q0  Q f )(1  ( NJ )a )b

(12)

This means that particles are accelerated


during the collision time W . The modified
&
&
macroscopic velocity u  'u enters the
calculation of the equilibrium distribution fi eq via
equation (6).

100
10
1
0.1

0.1

10

Shear strain rate [1/s]

Fig 2: Approximation of the data reported by


Kawakami et al. (1994) by equation (13)
with a = 2, b = 2.055, and N = 2.86.

BOUNDARY CONDITIONS

When applying the lattice Boltzmann


method the rheological model is expressed in
terms of W , i.e. W Wf  ( W0  Wf )(1  ( NJ )a )b .
The shear-thinning model was also used by
Artoli and Sequeira (2006) for computing
arterial blood flow by following the formulation
given below.
The momentum flux tensor in the
coordinates system (D,E) is obtained by

The only boundary condition implemented in


this study is the mid-plane bounce-back
condition. Distribution functions, arriving at a
rigid surface are simply thrown back in the
opposite direction. For a detailed description
see Succi (2001).
RHEOLOGICAL MODEL

One advantage of the lattice Boltzmann


method over classical Navier-Stokes solvers is

499

1)
3 (DE

f
0 i

(1) & &


eiDeiE

2U 't WcC (SDE )

proposal by Krner et al. (2003). This single


phase approach distinguishes between filled,
emptied and partially filled grid cells. Filled cells
are normal cells during calculation while empty
cells are ignored. The partially filled cells build a
closed interface between the two other types
and have to be handled in a different manner,
since distribution functions are not defined in
empty neighbor cells. They have to be
reconstructed before the collision step occurs.
For a detailed description of this procedure the
reader is referred to Krner et al. (2003).

(14)

where C is a lattice dependent constant (for


the two-dimensional model D2Q9 we have
C 1/ 3 ) and the constant Wc is set equal to 1.
The strain rate tensor is:


SDE

3
2U 'tWc

f
0 i

(1) & &


eiDeiE

(15)

The viscosity in shear-thinning fluids is


defined in terms of the second invariant of the
strain rate tensor S

S { J

2SDE : SDE

ILLUSTRATIVE EXAMPLES AND RESULTS

We solved using the procedure outlined


above two boundary value problems. The first
one refers to a slope subjected to an impulse
load in form of a cylindrical pressure wave in
the interior. The second example is a rigid wall
embedded in sloping ground whereby the flow
is induced by a series of cylindrical pressure
waves randomly distributed over the liquefiable
surface layer.
The simulation of collapse in a Bingham
fluid (slump test) was used for verification
purposes.
The set of data used for the numerical
simulations
are
given
in
so-called
dimensionless lattice values. The transfer to
physical quantities (denoted by star) is made by
the following equations:

(16)

Using symmetry of the strain rate tensor we


get the simple form for the strain rate in a ( x, y )
coordinates system

J

2
2
2
J c 0.5(Sxx
 Syy
)  Sxy

(17)

with
J c

3
2U 'tWc

(18)
(0)

we are now able to


Assuming fi(1) | fi  fi
calculate the strain rate during the collision step
with minimal computational effort.
As an extension of the standard CarreauYasuda model we introduce here a simple
retardation principle during the collision step. It
means that below a certain shear strain rate
level J lim the velocity of the particles is reduced
by a factor M .

&
u i*

&
M u
&

|
|

J  J lim
J t J lim

g*

'x
't

&
u*

& 'x
u
;
't

Q*

't
'x 2

The parameters of the rheological model are


selected to describe a typical viscous
behaviour. The use of higher viscosity values
(such as those depicted in Fig 2 for liquefied
sand) would merely require a shorter time step
(and accordingly a denser grid).
The data set used in the numerical
simulations is:
Parameters of the rheological model:
W f 0.60 ; W0 6 ; N 500 ; a 2.1;
b 2.1 ;
J lim = 0.001; M =0.3.
Grid geometry: 'x 'y 1
Grid size: 200 units in length (x); 100 units in
height (y)
Time step duration: 't 1

(19)

The motivation for this assumption is that it


leads to more plausible results when simulating
the erosive character of the flow process in
liquefied sands.
MODELING OF THE FREE SURFACE

For modeling the free surface we used the

500

Density of fluid U 1
Acceleration of gravity: g = 0.001
Sound propagation velocity in fluid cs 1
Slope angle: 55
The results are given in Fig 3 and Fig 4 as
snap shots of the velocity distribution within the
liquefied sand.
The computational time is 20 ms per time
step in a Pentium IV at 2.8 GHz.
A qualitative assessment of the forces
acting of the wall can be determined from the
distribution functions. Algorithms for simulating
open boundaries at both sides of the
computational domain have still to be
developed for this numerical method when
dealing with boundary value problems involving
free surfaces.

Krner, C., Pohl T., Rde, U., Threy, N. and Zeiser,


T. (2003) Parallel Lattice Boltzmann Methods for
CFD Applications, Lecture Notes in Computer
Science and Engineering, Springer, Berlin, Vol.
51, No. 2, pp. 439-465
Qian Y., d Humeres D., Lallemand P. (1992)
Lattice BGK for Navier-Stokes Equation,
Europhysics Letters 17, Vol. 6, pp. 479-484
Sato, T., Y. Moon and R. Uzuoka (2002) Unified
Analyses of Liquefaction and Flow Processes
of Inclined Ground Using Fluidal Elasto-Plastic
Model, J. of Geotechnical Engineering, JSCE,
No.722, III-61, pp.109-119.
Seed, R.B. and L.F. Harder, Jr. (1990) SPT-Based
Analysis of Cyclic Pore Pressure Generation
and Undrained Residual Strength, Proc. H.
Bolton Seed Memorial Symposium, Vol. 2, pp.
351-376.
Succi S. (2001) The Lattice Boltzmann Equation for
Fluid Dynamics and Beyond, Clarendon Press
Uzuoka, R., Yashima, A., Kawakami, T. and J.-M.
Konrad, J.-M. (1998) Fluid Dynamics Based
Prediction of Liquefaction Induced Lateral
Spreading, Computers and Geotechnics, Vol.
22, pp. 243-282.
Wolf-Gladrow, D.A. (2000) Lattice Gas Cellular
Automata and Lattice Boltzmann Models: An
Introduction, Lecture notes in mathematics No.
1725, Springer
Yu D., Mei E., Luo L.-S. and Shyy W.
(2003) Viscous Flow Computation with the
Method of Lattice Boltzmann Equation,
Progress in Aerospace Sciences, Vol. 39, pp.
329-367

CONCLUSIONS

The lattice Boltzmann method is a powerful


tool for a variety of fluid dynamics problems and
can be applied to predict the flow of liquefied
sand. A variety of rheological models can be
used within this formulation. Future work
includes the proper modeling of boundaries, the
optimization for higher viscosity values, and
finally the extension to describe more complex
material behavior.
REFERENCES
Artoli, A.M. and Sequeira, A. (2006) Mesoscopic
Simulations of Unsteady Shear-Thinning Flows,
International Conference on Computational
Science (2), Springer, pp.75-85
Bhatnagar P.L., Gross E.P. and Krook M.A. (1954)
A Model for Collision Processes in Gases, I.
Small Amplitude Processes in Charged and
Neutral One-Component Systems, Phys Rev,
Vol. 94, pp. 511-525
de Alba, P. and Ballestero, T.P. (2006) Residual
Strength after Liquefaction: A Rheological
Approach, Soil Dyn. Earthquake Eng, Vol. 26,
pp. 143-151
He, X. and Luo, L.-S. (1997) "A priori Derivation of
the Lattice Boltzmann Equation", Phys. Rev. E,
Vol. 55, pp. R6333-R6336
Kawakami, T., Suemasa, N., Hamada, M., Sato, H.
and. Katada, T. (1994) Experimental Study on
Mechanical Properties of Liquefied Sand, Proc.
Fifth US-Japan Workshop on Earthquake
Resistant Design of Lifeline Facilities and
Countermeasures against Soil Liquefaction.
Report NCEER-94-0026, pp. 285-299.

501

Fig 3: Collapsing slope: Velocity plot at different time steps

502

Fig 4: Rigid wall in liquefied layer: Velocity plot at different time steps

503

On the Liquefaction Resistance of Silty Sands


A. I. Papadopoulou1, T. M. Tika2
1

Post Graduate Student, Aristotle University of Thessaloniki, Greece


2
Associate Professor, Aristotle University of Thessaloniki, Greece

Abstract
The paper presents results from a laboratory investigation into the liquefaction and critical
state characteristics of silty sands. The critical state line and liquefaction resistance of
natural silty sands and artificial mixtures of a clean sand with a non-plastic silt at various
contents were studied. It is shown that the location of the critical state line is different for
each tested soil. At low stresses, the critical state lines are nearly parallel and horizontal
and they tend to steepen and converge with increasing mean effective stess. The
liquefaction resistance of the tested soils decreased with increasing void ratio at constant
mean effective stress. The study showed that there is a unique relationship between the
liquefaction resistance and relative state parameter for each soil.

is above the critical state line, the sand has the


tendency to contract upon shearing, whereas
when its state is below the critical state line, it
has a tendency to dilate. Various normalized
parameters
have
been
proposed
to
characterize the difference between the actual
state and the critical state line. Been and
Jefferies (1985) have quantified the distance of
the current state from the critical state line by
means of a state parameter, (or ), which is
the difference in void ratios, e, between the
current state and the critical state line at the
current mean effective stress, p, Fig 1.
Recently, in an attempt to reevaluate the
effect of overburden stress, considered in the
semi-empirical field-based procedures for
evaluating liquefaction potential, the liquefaction
resistance of clean sands during both
monotonic and cyclic loading has been
correlated with the distance of their initial state
from the critical state line (Pillai & Muhunthan,
2001, Boulanger, 2003 and Idriss & Boulanger,
2004).
The purpose of the work presented in this
paper is to study the liquefaction resistance and
critical state characteristics of silty sands.

INTRODUCTION

Liquefaction of sandy soils under monotonic


and cyclic loading conditions is considered to
be one of the major causes of failure of earth
structures and foundations. Field observations
from failures due to liquefaction have shown,
that both clean sands and sands containing
fines (%<75m) are susceptible to liquefaction.
Since the pioneering works of Casagrande
(1936) and Schofield & Wroth (1968), the
critical / steady state concept for sands has
increasingly been used as a fundamental state
to characterize the strength and deformation
properties of sands in limit equilibrium analyses.
Under undrained conditions, the critical state is
reached when the pore pressure and the shear
stress remain constant during continued
deformation. At this stage, there is a unique
relationship among void ratio, ecs, mean
effective stress, pcs, and shear strength, qcs,
expressed by the critical state line (CSL) in the
void ratio versus mean effective stress plane.
According to the critical state concept, the
behaviour of a sand depends not only on
density, but also on stress level. The true state
of a sand is described by the location of its
current state of stress and volume relative to
the critical state line. When the state of a sand

504

by (Ladd, 1978). This sample preparation


method was preferred, as it produces uniform
specimens at various densities. Saturation was
achieved by means of the carbon dioxide (CO2)
method. In all the tests the parameter of pore
water pressure, B=u/, had values from 0.95
to 1.00. The specimens were then isotropically
consolidated to the desired effective stress.
Tests were conducted at various densities and
mean effective stresses.
In the monotonic tests, the specimens were
subjected to undrained compression at a
constant rate of axial deformation varying from
0.05 to 0.1mm/min, whereas in the cyclic tests,
a sinusoidally varying axial stress (d) was
applied at a frequency of f=0.1Hz, under
undrained conditions. In this work, liquefaction
resistance or cyclic resistance ratio, CRR15, is
defined as the cyclic stress ratio, CSR=d/2o,
required to cause double amplitude axial strain,
DA5% at 15 cycles of loading.

TESTED MATERIALS

The materials used in the testing


programme were natural silty sands and
artificial mixtures of sand and silt. Two natural
silty sands (Dzce-1, Dzce-2) were retrieved
from liquefied sites in Dzce, Turkey, during the
1999 earthquake of magnitude Mw=7.1 and the
other was retrieved from the foundation of the
Rimnio embankment bridge in Greece, which
failed during the 1995 Kozani earthquake of
magnitude Ms=6.6 (Tika & Pitilakis, 1999). The
artificial mixtures were prepared by mixing a
clean quartz sand (S) with well-rounded grains
with a non-plastic silt (F) at contents of 15, 25,
35 and 40% of total dry mass (noted as SF15,
SF25, SF35 and SF40 respectively). The grain
size distributions and physical properties of the
tested soils are presented in Fig 2 and Table 1,
respectively.
Although the ASTM (D 4243 & D 4254) test
methods for the determination of minimum and
maximum void ratios are applicable to soils that
may contain up to 15%, by dry mass, of soil
particles passing a No. 200 (75-m) sieve,
provided
they
have
cohesionless
characteristics, both these methods were used
in this work in conjunction with others in order to
get a consistent value. In particular, for the
determination of the minimum void ratio both
the vibratory table (ASTM D 4253, Method 2A)
and the Standard Proctor test methods were
used. The maximum void ratio was determined
in accordance with the ASTM (D 4254, Method
C) test method and also by pouring dry material
in a mould three times and considering the
average as the maximum value. Fig 3 shows
the minimum and maximum void ratios of the
tested soils.

TEST RESULTS AND DISCUSSION

Several studies have considered sands


containing fines as consisting of two matrices,
the sand grains matrix and the fine matrix and
analysed the interaction with each other (Vaid,
1994, Thevanayagam et al., 2000, Polito &
Martin, 2001, Xenaki & Athanasopoulos, 2003
and Yang et al., 2004). At low fines content, the
fines are mainly contained within the
intergranular voids of sand grains and may not
actively participate in transferring or sustaining
any forces. However, with increasing fines
content up to the threshold value, they not only
fill the intergranular space, but they begin to
displace the sand grains contacts. If the fines
content exceeds the threshold value, the
behaviour of the mixture is primarily governed
by the contacts between the fines and the sand
grains float within the fines matrix.
Fig 4 shows the variation of liquefaction
resistance with void ratio at a mean effective
stress of 100kPa. A decrease of liquefaction
resistance with increasing void ratio is observed
for each tested soil. For the artificial mixtures, it
is also indicated that at a given void ratio,
liquefaction resistance initially decreases with
increasing fines content up to 35% and
increases thereafter with further increasing
fines content. This threshold fines content
depends on the characteristics of sand and silt
and the mean effective stress level

TESTING PROGRAMME

The testing programme consisted of


monotonic and cyclic triaxial tests for the
determination of critical state line and
liquefaction resistance of soils, respectively.
Both types of tests were performed using a
cyclic triaxial apparatus (MTS Systems
Corporation, U.S.A). Its principles of operation
are given in detail in Papadopoulou (2007).
The specimens (height / diameter 100mm
/ 50mm) were formed by moist tamping at a
water content varying between 4 and 35%,
using the undercompaction method, introduced

505

Fig 1 Definition of state parameter (Been and Jefferies, 1985)

U.S.A. ASTM D422


FINE

SILT
No. 200

100

MEDIUM COARSE

COBBLES

GRAVELS

SAND
No. 100

No. 50

No. 40

3 inch

0
Sand
10
Silt
Dzce-1
20
Dzce-2
Rimnio 30

90

Percent passing: %

80
70

Liquefiable
Soils

60

40

50

50

Probable
Liquefiable Soils

40

60

30

70

20

80

10

90

0
0.001

Percent retained: %

CLAY

100
0.010

0.100

1.000

10.000

Particle size: mm

Fig 2: Grain size distributions of tested soils

506

100.000

1000.000

Table 1: Physical properties of tested soils


Soils

Gs

emax

emin

D50
(mm)

Cu

fc (%)
(<75m)

CF (%)
(<5m)

Classification
USCS

Natural :
Dzce-1
Dzce-2
Rimnio

2.732
2.741
2.667

0.724
1.155
1.059

0.366
0.521
0.419

0.38
0.19
0.10

9.7
12.5
5.3

13
23
37

5.0
3.0

SM
SM
SM

Artificial :
Sand
Silt

2.649
2.663

0.841
1.663

0.582
0.658

0.30
0.02

1.3
7.5

0
100

0
12.0

SP
ML

Void ratio, e

1.6

1.2

0.8

0.4

0
0

10

20

30

40

50

60

70

80

90

fines content, fc: %

Artficial mixtures
Dzce-1
Dzce-2
Rimnio

emax

emin

Fig 3: Variation of maximum, emax, and minimum, emin, void ratios with fines content, fc

507

100

0.6

(a)
0.4
CRR15
0.2

0
0.4

0.5

0.6

0.7

0.8

0.9

void ratio, e
0.6

(b)
0.4
CRR15
0.2

0
0.4

0.5

0.6

0.7

0.8

void ratio, e

Artificial soils
Sand (S)
SF15
SF25
SF35
SF40
Silt (F)

Natural soils
Dzce-1
Dzce-2
Rimnio

Fig 4: Variation of cyclic resistance ratio, CRR15, with void ratio, e, for (a) natural silty sands and (b) artificial
mixtures, at a mean effective stress of =100kPa

508

(Papadopoulou & Tika, 2007, Papadopoulou,


2007).
The critical state lines of the tested soils
are shown in Fig 5. It is indicated that the
location of the critical state line is different for
each soil. At low stresses, the critical state
lines are nearly parallel and have a small
inclination. With increasing mean effective
stess level, however, they steepen and
converge at stesses above 1000kPa
approximately. Moreover, for the artificial
mixtures the critical state lines move
downwards with increasing fines content up to
35%, and then upwards. The critical state
lines at fines content of 35% and 40% are
very close, whereas the critical states lines of
the clean sand and silt practically coincide.
The fines content at which the location of
critical state line changes is of the same order
with the threshold fines content, stated above,
and it is the fines content at which the mixture
starts to behave as silt with sand inclusions.
The liquefaction resistance of the tested
soils is related to the position of their initial
state relative to their critical state through a
relative state parameter, R, in Fig 6. This
parameter was obtained by normalizing the
value of state parameter, , by the difference
in the maximum and minimum void ratios
(emax-emin) of each tested soil and provides a
convenient way to consider the combined
influence of density and mean effective stress
on the behaviour of soils (Konrad, 1988,
Boulanger, 2003). As shown in the above
figure, there is a unique relationship between
liquefaction resistance and relative state
parameter for each tested soil. A good
agreement of the results from Dzce-1 and
Dzce-2 and the corresponding from the
artificial mixtures with similar fines content is
observed. For Rimnio silty sand higher values
of liquefaction resistance at a given value of
R was determined, as compared with the
artificial mixtures with 35% and 40% fines
content. This difference may be attributed to
differences in gradation and particle
roundness and angularity. It should be also
noted that the in-situ relationship between
liquefaction resistance and relative state
parameter may be different from that
determined in laboratory due to the effect of
stress history, structure (fabric and bonding)
and ageing.

CONCLUSIONS

The following conclusions can be drawn


regarding the critical state and liquefaction
characteristics of the tested soils:
i). Liquefaction resistance decreases with
increasing void ratio at constant mean
effective stress. For the artificial mixtures
tested at a given void ratio, liquefaction
resistance decreases initially up to a threshold
fines content of 35% and increases thereafter
with further increasing fines content.
ii). The location of the critical state line is
different for each soil. At low stresses, the
critical state lines are nearly parallel and have
a small inclination. With increasing mean
effective stess level, however, they steepen
and converge at stresses above 1000kPa
approximately. For the artificial mixtures
tested, the critical state lines move
downwards with increasing fines content up to
35%, and then upwards.
iii). A
unique
relationship
between
liquefaction resistance and relative state
parameter is observed for each tested soil.
Differences between the natural and artificial
soils may be attributed to differences in
gradation and particle roundness and
angularity.
AKNOWLEDGEMENTS

The authors would like to express their


acknowledgements to the State Scholarships
Foundation of Greece and the General
Secretary of Research and Technology,
Greece, for the financial support of the first
author. The tests on the soils from Dzce
were performed within the Marmara
Earthquake Rehabilitation Project (MERP).
REFERENCES
Been, K., and Jefferies, M.G. (1985) A state
parameter for sands, Gotechnique, Vol. 41,
No. 3, pp. 365-381
Boulanger, R.W. (2003) High overburden stress
effects in liquefaction analyses, Journal of
Geotechnical
and
Geoenvironmental
Engineering, ASCE, Vol. 129, No. 12, pp. 10711082
Casagrande, A. (1936) Characteristics of
cohesionless soil affecting the stability of slopes

509

(a)
0.8

ecs 0.6

0.4

0.2

10

100

1000

10000

log p'cs: kPa


1

(b)
0.8

ecs 0.6

0.4

0.2

10

100

1000

10000

log p'cs: kPa


Artificial soils
Sand (S)
SF15
SF25
SF35
SF40
Silt (F)

Natural soils
Dzce-1
Dzce-2
Rimnio

Fig 5: Critical state lines of tested soils, (a) natural silty sands and (b) artificial mixtures

510

0.6

0.6

SF15
Dzce-1

Sand (S)

(a)

0.4

(b)

0.4

CRR 15
0.2

0.2

0
-0.8

-0.4

0.4

0.8

0
-0.8

-0.4

0.4

0.6

0.6

SF35
SF40
Silt (F)
Rimnio

SF25
Dzce-2

(c)

0.4

0.8

0.4

(d)

CRR 15
0.2

0.2

0
-0.8

-0.4

0.4

0.8

0
-0.8

-0.4

0.4

0.8

Relative state parameter, R


Artificial soils
Sand (S)
SF15
SF25
SF35
SF40
Silt (F)

Natural soils
Dzce-1
Dzce-2
Rimnio

Fig 6: Variation of cyclic resistance ratio, CRR15, with relative state parameter, R, at constant mean effective
stress, =100kPa

511

Xenaki, V.C., and Athanasopoulos, G.A. (2003)


Liquefaction resistance of sand-silt mixtures:
an experimental investigation of the effect of
fines, Soil Dynamics and Earthquake
Engineering, Vol. 23, pp. 183-194
Yang, S., Sandven, R., and Grande, L. (2004)
Instabillity of loose sand-silt mixtures,
International
Proceedings
of
the
11th
Conference on Soil Dynamics and Earthquake
Engineering-3rd International Conference on
Earthquake
Geotechnical
Engineering,
Berkeley, California, U.S.A.

and earth fills, Journal of Boston Society of


Civil Engineering, pp. 13-32
Idriss, I.M., and Boulanger, R.W. (2004) Semiempirical procedures for evaluating liquefaction
potential during earthquakes, Proceedings of
the 11th International Conference on Soil
Dynamics and Earthquake Engineering- 3rd
International Conference on Earthquake
Geotechnical Engineering
Konrad, J.M. (1988) Interpretation of flat plate
dilatometer tests in sand in terms of the state
parameter, Gotechnique, Vol. 38, No. 2, pp.
263-277
Ladd, R.S. (1978) Preparing test specimens using
undercompaction,
Geotechnical
Testing
Journal, GTJODJ, Vol. 1, pp. 16-23
Papadopoulou, A.I., Tika T.M. (2007) The effect of
non-plastic fines on the liquefaction resistance
of sands, Proceedings of the 4h International
Conference on Earthquake Geotechnical
Engneering, Thessaloniki, Paper No. 1414
Papadopoulou, A.I. (2007) Liquefaction resistance
of sands with non-plastic fines, Ph.D Thesis,
Aristotle University of Thessaloniki (under
preparation)
Pillai, V.S., and Muhunthan, B. (2001) A review of
the influence of initial static shear (Ka) and
confining stress (K) on failure mechanisms and
earthquake liquefaction of soils, Proceedings
of the 4h International Conference on Recent
Advances
in
Geotechnical
Earthquake
Engineering and Soil Dynamics, San Diego,
California, Paper No. 1.51
Polito, C.P., and Martin, J.R. (2001) Effects of
nonplastic fines on the liquefaction resistance
of sands, Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, Vol.
127, No. 5, pp. 408-415
Schofield, A.N., and Wroth, P. (1968) Critical State
Soil Mechanics, Mc-Graw Hill, London
Thevanayagam, S., Fiorillo, M., and Liang, J.
(2000) Effect of non-plastic fines on undrained
cyclic strength of silty sands, Soil Dynamics
and Liquefaction, pp. 77-91
Thevanayagam, S., and Mohan, S. (2000)
Intergranular state variables and stress-strain
behaviour of silty sands, Gotechnique, Vol.
50, No. 1, pp. 1-23
Tika, T., and Pitilakis, K. (1999) Performance of
Rimnio bridge embankment during the 1995
Kozani-Grevena earthquake, Proceedings of
the 12th European Conference on Soil
Mechanics and Geotechnical Engineering, The
Netherlands, Vol.2, pp. 857-862
Vaid, V.P. (1994) Liquefaction of silty soils,
Ground Failures Under Seismic Conditions,
Geotechnical Special Publication, No. 44,
ASCE, pp. 1-16

512

Effect of Countermeasures for Pile Foundation under Lateral Flow


Caused by Liquefaction
T. Matsuda and K. Sato
Technical Research Institute, Obayashi Corporation, JAPAN

Abstract
The authors have developed countermeasures for pile foundation during lateral flow caused
by ground liquefaction. Three countermeasures had been designed with taking account for
reducing the flow forces of unliquefiable layers above the liquefiable layers. The basic ideas
of these countermeasures are reduction of acting forces to footing structure of piles. Effects
of these countermeasures were verified by a dynamic centrifuge experiment. The result of
the experiment shows that the effect for residual deformation of the footing and total
stresses of piles were significantly large. The residual lateral deformation of footing with the
streamline facing as a countermeasure was 70% smaller than that of the case without
countermeasures. Also, the maximum stresses of the piles with countermeasures were kept
elastic conditions, while those of the case without any countermeasures exceeded the yield
stresses.

INTRODUCTION
Unliquefiable layer

Damage of pile foundations caused by


ground lateral flow has been observed along
backfill of quay wall during and after strong
earthquakes. It is thought that acting forces to
footings and piles under unliquefiable surface
layers are larger than those in the liquefiable
lower
layers,
which
means
that
countermeasures against the lateral ground
flows to piles are more effective in the case of
implementation to upper unliquefiable layers.
The
authors
have
developed
three
countermeasures based on that viewpoint.
These countermeasures are earth retaining
walls in front of the footings, drain piles from
liquefiable layers to unliquefiable layers and
streamlined shield facing implemented in front
of the footing.
A dynamic centrifuge test for these
countermeasures was performed to verify the
effect of these methods. This paper describes
the result of the test and the mechanism of the
prevention for damage on piles.

Footing

Caisson

Pore water
Liquefiable layer
[ Drain piles method ]

Earth retaining wall


Ground flow

Trench

[ Earth retaining wall method ]

Ground flow

[ Stream shield block method ]

Figure 1: Countermeasures for ground lateral flow

513

COUNTERMEASURES

Three countermeasures were verified by the


centrifuge test as shown in figure 1.
Drain piles method

The permeability of gravel drain piles around


footing should be smaller than that of
surrounding liquefiable ground. These gravel
drain piles are installed from the bottom of
liquefiable layers to the midpoint of upper
unliquefiable layers. The gravel drain piles have
effects not only on the prevention to liquefaction
of lower liquefiable layers, but also decreasing
the stiffness of the upper unsaturated layers
associated with the footing deformation.
Namely, the excess pore water pressures built
up in the lower liquefiable layers are transmitted
to the upper unsaturated layer by the drain piles,
the stiffness of the upper unsaturated layer is
deteriorated by the excess pore water
pressures propagated from the lower layers.
Earth retaining wall method

In this method, an earth retaining wall is


installed in front of the footing foundation. The
direction of ground lateral flow should be
determined, so that the retaining wall can
directly block lateral ground flow from upper
side.
Streamlined shield block method

with a steel footing.


The concrete caisson was designed to resist
the earth pressure at rest by the friction
between bottom of the caisson and the bearing
soil layer. All dimensions of the centrifuge
model are described by prototype scale and
model scales are described in parentheses.
The backfill layer consisted of upper
unsaturated sand layer of 3m thickness (6cm in
model scale) and lower saturated sand layer of
12m (24cm). The material of sand layer was
uniform silica sand of No.8 Japanese standard,
and physical properties of used sand are
tabulated in Table 1. The model ground was
manufactured by air dispersion method, and the
density of sand layers was tuned by dropping
height and application amount. The de-aired
normal water was used for the pore water, so
that the coefficient of permeability of prototype
ground should be 50 times of model ground.
Before injections of degas water, carbon
dioxide was filled in the pore. The physical
properties of the model ground are tabulated in
Table 2. The relative density of the upper
unsaturated layer was 58% and that on the
lower saturated layer was 7% at 1g fields. The
relative density of the lower layer was
consolidated by centrifugal acceleration, the
final density reached about 20% at 50g.

Table 1: Material properties of sand

In this method, an acutely angle facing of


the footing foundation is casted. The acute
angle face can disperse soils from lateral
direction to slant direction or vertical direction.

Grain density

DYNAMIC CENTRIFUGE TEST

Gs (t/m3)

2.650

Max. Grain size Dmax (mm)

0.25

Max. void ratio emax

1.403

Min. void ratio emin

0.705

Models

Cross and plan sections of the test models


are shown in Figure 2. A rigid soil container on
the shaking table of the centrifuge machine
were divided by 4 cells. The centrifugal
acceleration was 50g, and four models (one
without
countermeasures
and
three
countermeasures) were shaken simultaneously.
Each cell (soil container) has 47.5m long by
20m widths in prototype scale.
Each centrifuge model consisted of steel
caisson facing to waterfront side, backfills as
liquefiable layers and group piles foundation

Table 2: Material properties of models

Density
(t/m3)
Unsaturat
ed
Saturated
Caisson

514

1.33
(Dry)
1.70
2.16

Relative
density
(%)

Permeability
(m/Sec)

58

8.5x10-3

Plan View
(1) Without-countermeasures

(2) Drain piles method

Caisson

Drain piles

5cm (2.5m)

Steel footing

(3) Earth retaining


wall method

40cm
(20m)

Gravel mat

24cm (12m)

5cm (2.5m)

(4) Streamlined shield block method


40cm
(20m)

Earth retaining wall

95cm (47.5m)

95cm (47.5m)
50cm (25m)

10cm (5m) 10cm (5m) 15cm (7.5m)


10cm (5m)

190cm (95m)

Section View
(2) Drain piles method

(1) Without-countermeasures
AH-12

Unliquefiable layer (Dr=58%)


Liquefiable layer (Dr=7%)
WP-01
AH-11

LH-02

LH-01
WP-24
ST-11
ST-12
Stainless pile

WP-21

ST-13
a

Foundation layer
(Soil mortar)

WP-25
WP-26

ST-21

WP-22

ST-22

WP-23

ST-23

b
pile

(3) Earth retaining


wall method

(4) Streamlined shield block method


LH-03

6cm (3m)
Earth retaining wall
24cm
(12m)

LH-04
ST-31

WP-31
ST-32

30cm
(15m)

WP-41

ST-33

12cm
(6m)

ST-41
ST-42

12cm
(6m)

ST-43

20cm
(10m)
Accelerometers

Pore water pressure meters

Strain gauges
Shaking

Figure 2: Dynamic centrifuge models

515

Displacement gauges

Input motion

The coefficient of permeability of prototype


ground was computed as 8.5x10-3m/sec in
prototype ground. The bearing ground layer
was installed under saturated sandy ground
using soil mortar of 10m (20cm) thickness.
The caisson structure consisted of steel
footing and 4 piles (SUS304) of 1m (2cm) in
diameter and 25mm (0.5mm) in thickness. 4
piles were supported into the bearing layer in
9m (18cm) depth.
In the case of the drain piles method, gravel
drain piles consisted of coarse-grained soils of
5cm (1mm) in diameter. 32 gravel drain piles
were installed around the footing by 2.5m (5cm)
spacing. Each gravel pile was connected to
gravel flat mat in the upper unsaturated layer to
dissipate the pore water pressure into those
layers.
In the case of the earth retaining wall
method, the deep trench of 2.5m (5cm) in width
was installed in front of the footing at the upper
flowing side. This trench was maintained by the
steel angle of 3m by 3m (6cm by 6cm).
In the case of the streamlined shield block
method, wedge shape of footing side face had
an angle of 90 degree. The crest of wedge as
shield block was located in the middle of the
footing.

The maximum amplitude of the input motion


at the shaking table was set up to 600Gal (30g).
20 waves of sinusoidal wave in 1.2Hz (60Hz)
were subjected. The observed time history at
the bottom of the saturated sand layer is shown
in Figure 3. Under the influence of liquefaction
of sandy layers, the input motion to the sandy
layers was degraded and decentered.

1000
(1) Without-contermeasures (AH-11)

500
0
-500
-1000
0

10

15

20

RESULTS

Lateral displacement

An average residual lateral displacement


was 4.03m (8cm) as tabulated in Table 3. It is
thought that the differences among the test
cases were caused not only from fluctuation of
manufacturing
models
but
also
from
performances of the countermeasures.

Measurements

Measuring instrumentations were for


accelerations (AH), pore water pressures (WP),
displacements of the footings and caissons
(LH) and strains of the piles (ST). Two strain
gages were installed on the upstream side and
downstream side of ground lateral flow in the
part of pile head, midst part and boundary part
to the bearing layer.

Table 3: Maximum lateral displacement of caisson

Case
1. Without-countermeasures
2. Drain piles method
3. Earth retaining wall method
4. Streamlined shield block method

(sec) 25

Figure 3: Input motion at the bottom of saturated


liquefiable layer in prototype scale

Residual Lateral Displacement of Caisson


3.3 m
4.0 m
3.8 m
5.0 m

516

Time histories of lateral displacement at the


top of the footing during shaking are shown in
Figure 4. The minus (-) sign indicates direction
of lateral ground flow. The maximum lateral
displacement is about 1m in the case of without
countermeasures, which is almost 1/4 of that of
the caisson. While the input motion was
subjected from 2 second to 19 second, the
lateral displacement progressed just after
beginning
of
shaking.
The
maximum
displacement was observed during the first half
of shaking, and during the last half of shaking,
the lateral displacement slightly decreased.
This rebound of the lateral displacement would
be affected by the elastic component of piles.
The largest lateral displacement among test
cases was observed at the case of withoutcountermeasures. The lateral displacements in
case of the earth retaining wall and the
streamlined
shield
block
method
are
significantly decreased from that of withoutcountermeasures. In the case of withoutcountermeasures the foundation suffered
lateral forces not only from the saturated layer
but also from upper unsaturated layer, while in
the case of the earth retaining wall the footing
did not suffer earth pressures from upper
unsaturated
layers.
The
maximum
displacement of the earth retaining wall was
about 50% of that in case of withoutcountermeasures,
which
indicates
that
decrementation of flow forces along the
unliquefiable layers is significantly effective for
damage of foundations.
The
vibration
components
on
the
displacement time histories were between 15
and 20 cm irrespective of the test cases. This
result indicates that the countermeasures
mainly act for mitigation of lateral residual
deformations.
Figure 5 shows the residual displacements
at the top of the footings. Those of any
countermeasures were smaller than that of
without-countermeasures on 75cm. The
minimum displacement was observed at the
streamlined shield block method, which was
12% of without countermeasures.

20
0
-20
-40
-60
-80
-100
0

(1) Without-countermeasures (LH-01)

20
0
-20
-40
-60
-80
-100
0

10

15

20

(sec) 25

(2) Drain piles method (LH-02)


Malfunction

20
0
-20
-40
-60
-80
-100
0

10

15

20

(sec) 25

(3) Earth retaining wall method (LH-03)

10

15

20

(sec) 25

20
0
-20
Malfunction
-40
-60
-80 (4) Streamlined shield block method (LH-04)
-100
0
5
10
15
20 (sec) 25

Figure 4: Time histories of lateral displacement

90
75
60

(1) Without-countermeasures
(2) Drain piles
method

45
30

(3) Earth retaining


wall method

15
0
-15

Figure

517

5:

(4) Streamlined shield


block method

Comparison of residual lateral


displacement of footings

Pore water pressures

2
1.5
1
0.5
0
-0.5
0

Figure 6 shows time histories of excess


pore water pressure ratio in case of the drain
piles method. The maximum excess pore water
pressure ratio at WP-21 and WP-22 which were
located at the outside of drain piles reached 1.0,
which means the free field reached liquefaction.
Meanwhile WP-23 that was located around
drain piles did not reach 1.0 (except vibration
components), which indicated that drain piles
were effective for non-liquefaction. However,
the excess pore water pressure ratio at WP-23
was built up to 0.5, it was thought that the
stiffness of the ground below the footing
became about 75% of the initial stiffness. In this
state, it was impossible to prevent the lateral
flows toward the caisson.
As the result of liquefaction behind the
caisson, the residual lateral displacement of the
drain piles method occurred about 40cm, which
was about 55% of the without-countermeasures.
It was shown that the drain piles method can
reduce the residual displacement, however it
was not enough to prevent the lateral flows.
Regarding the excess pore water pressure
in the unsaturated layers such as WP-24 and
WP-25, the excess pore water pressures
slightly built up at the place far from drain piles
after the end of shaking of around 25 second.
However, the excess pore water pressure in the
gravel mat such as WP-26 is rapidly built up
during shaking (after 7 second), eventually the
excess pore water pressure ratio exceeded
over 3. It is indicated that drain piles are
effective for feeding pressurized water that lead
to decrease the stiffness of unsaturated layers.
In this centrifuge test, the lateral residual
displacement had already reached the
maximum displacement as described later
using Figure 7, which was not able to reduce
the deformation effectively.
Figure 7 shows the relationship between
lateral displacement of the footing and excess
pore water pressure (residual component) at
the saturated layer of the backfill ground. Both
without countermeasures and the streamlined
shield block method the footing moved to the
caisson in proportion to the excess pore water
pressure by the time when the excess pore
water pressure ratio built up to about 0.2. It is to
be noted that both two cases the residual
lateral displacement reached 50% of the
maximum displacement when the excess pore

2
1.5
1
0.5
0
-0.5
0

2
1.5
1
0.5
0
-0.5
0

2
1.5
1
0.5
0
-0.5
0

2
1.5
1
0.5
0
-0.5
0

4
3
2
1
0
-1
0

(2) Drain piles method


(WP-21)

10

15

20

25

30

35(sec)40

(2) Drain piles method


(WP-22)

10

15

20

25

30

35(sec)40

(2) Drain piles method


(WP-23)

10

15

20

25

30

35(sec)40

(2) Drain piles method


(WP-24)

10

15

20

25

30

35(sec)40

(2) Drain piles method


(WP-25)

10

15

20

25

30

35(sec)40

(2) Drain piles method


(WP-26)

10

15

20

25

30

35(sec)40

Figure 6: Time histories of pore water pressures

water pressure ratio built up to about 0.7. When


the pore water pressure ratio became between
0.7 and 0.9, the residual displacement
extremely progressed up to the maximum
displacement, eventually rebound deformation
is initiated.

518

Especially,
the
rebound
of
lateral
displacement of the streamlined shield was
significant. It was shown that the streamlined
shield makes the footing to set back to the
original position. The residual displacement
was slightly progressed toward the caisson
during dissipation of the pore water pressure.
Difference between without-countermeasures
and the streamlined shield block method is
obviously observed when the pore water
pressure ratio built up over 0.2, especially over
0.7.

4000 (1) Without-countermeasures


ST-11
2000 y
0
ST-12
-2000 y
-4000
ST-13
-6000
0
5
10
15
20 (sec) 25

4000 (2) Drain piles method


2000 y
0
-2000 y
-4000
-6000
0
5
10

(4) Streamlined shield


block method

1
0.8

0.4

(1) Withoutcountermeasures

0.2
0
-0.2
20

Figure 7:

0
-20
-40
-60
-80
Lateral displacement
of footings (cm)
(cm)

ST-22

ST-23
15

20 (sec) 25

4000 (3) Earth retaining wall method


ST-31
2000 y
0
-2000 y
ST-33
-4000
-6000
0
5
10
15

0.6

ST-21

ST-32

20 (sec) 25

4000 (4) Streamlined shield block method


ST-31
2000 y
0
-2000 y
ST-33
-4000
-6000
0
5
10
15

-100

Relationship between lateral


displacement and excess pore
water pressure

ST-32

20 (sec) 25

Figure 8: Time histories of bending strains

Strains of piles

Time histories of bending strains of piles


together with the yield strain of 1.4x10-3 are
shown in figure 8. In case of the withoutcountermeasures
the
maximum
strains
exceeded the yield strain at both pile head and
lower boundary. On the other hand, in cases of
both the earth retaining wall method and the
streamlined shield block method the maximum
strains of all portion of the pile were kept within
elastic condition. In case of the drain piles
method the maximum strain did not exceed the
yield strain except one at the boundary.
Figure 9 shows the comparison of residual
bending strains ratio (Normalized to without
countermeasures) among all test cases.
Regarding to the strains at the pile head, any
countermeasures can reduce the residual
bending strain at least 60% of without
countermeasures. Especially both the earth

2000

r : Proportion to the without-countermeasures case


Pile
top

r = 1.00
1000

r = 0.57

r = 0.36

-1000
()

r = 0.05
r = -0.04
(4) Streamlined shield
block method
(3) Earth retaining
wall method

r = 0.18

-2000
-3000
-4000

r = 0.94
r = 1.00

Boundary of
liquefable
layer

and
foundation

(2) Drain piles method


(1)
Without-countermeasures

Figure 9:

519

Comparison on the maximum


bending strain

retaining wall method and the streamlined


shield block method can reduce bending strains
more than 20% of without countermeasures.

REFERENCES
Higuchi, S. and Matsuda, T. (2002): Characteristics
of the External Forces acting on a Pile during
Liquefaction-induced Lateral Flow of the Ground,
Proceedings of the Eighth U.S.-Japan Workshop
on Earthquake Resistant Design of Lifeline
Facilities
and
Countermeasures
Against
Liquefaction, pp.497-505

CONCLUSIONS

The dynamic centrifuge test on three


countermeasures against lateral ground flow
caused by liquefaction was carried out. These
countermeasures are designed with reducing
the external forces to the footings. The earth
retaining wall method and the streamlined
shield block method are directly able to reduce
the lateral flow force to the footings. As a result
of the tests, the streamlined shield block
method indicated remarkable decrementation of
residual lateral displacement and bending strain
of piles.
It is noted that the prevention mechanism of
the streamlined shield were not only to reduce
the lateral forces due to the ground flow caused
by liquefaction but also to set back to the
original position. This countermeasure does not
require any soil improvement to prevent ground
liquefaction, so that the cost performance of
construction would be higher than other
countermeasures.

Higuchi, S. and Matsuda, T. (2002): Effects of


Liquefaction-induced Lateral Flow of the Ground
against a Pile Foundation, Proceedings of
International Conference of Physical Modeling in
Geotechnics, pp.63-68
Jang, J.H., Hirao, A., Kurita, M. and Hamada, M.
(2002): An experimental study on external forces
from flowing liquefied on foundations piles,
Proceedings of the Eighth U.S.-Japan Workshop
on Earthquake Resistant Design of Lifeline
Facilities
and
Countermeasures
Against
Liquefaction, pp.529-540
Matsuda, T. and Higuchi, S. (2002): Development of
the Large Geotechnical Centrifuge and Shaking
Table of Obayashi, Proceedings of International
Conference of Physical Modeling in Geotechnics,
pp.63-68
Ricardo, R., Abdoun, T.H and Dobry, R (2000):
Effect of lateral stiffness of superstructure on
bending moments of pile foundation due to
liquefaction-induced
lateral
spreading,
Proceedings of the 12th World Conference on
Earthquake Engineering, 0902

520

EFFECT OF DURATION OF EARTHQUAKE


ON ONSET OF LIQUEFACTION
Nozomu YOSHIDA
Tohoku Gakuin University, Japan

Abstract
The effect of long duration earthquake, the Tonankai earthquake, which is an ocean trench
type earthquake and is expected to come in near future, is investigated from the point of
view of liquefaction. A synthesized earthquake motion for this earthquake at the Amagasaki
site is used. The earthquake motion observed during the 1995 Hyogoken-nambu
earthquake is also used for comparison purpose. Since liquefaction is not expected at the
site, Port Island vertical array site is used in the analysis. Liquefaction does not occur
against Hyogoken-nambu earthquake wave, but it occurs under the future Tonankai
earthquake. Detailed investigation indicates that huge numbers of cycles is a key factor to
cause liquefaction although shear strain amplitude is smaller.

INTRODUCTION

motion with long duration hit the Tomakomai


Port and an oil storage tank fired because of
resonance with long period earthquake motion.
Number of cycles looks much larger than those
discussed in the preceding.
In Japan, a bigger earthquake than the
Tokachi-Oki earthquake is expected in the
Pacific Ocean side. They are ocean trench type
earthquakes named Tokai, Tonankai, and
Nankai earthquakes. Figure 1 (Central Disaster
Prevention Council, 2005) shows fault zone of
the Tonankai and Nankai earthquakes and
expected JMA seismic intensity distribution

Onset of soil liquefaction is frequently


identified by, so called, FL method, a method
based on liquefaction resistant factor. In this
method, expected maximum shear stress is
compared with liquefaction strength. Here,
liquefaction strength is evaluated based on
constant stress amplitude test, and, therefore, it
cannot be compared with maximum shear
stress directly. Various factors are taken into
consideration in order to convert the
liquefaction strength in the laboratory into the
liquefaction strength during earthquake. They
are effects of sampling and handling, effect of
coefficient of earth pressure at rest, irregular
nature of earthquake motion, and multidirectional shaking (Iwasaki et al., 1978).
Among them, the effect of irregular nature of
earthquake motion is estimated to be between
1/0.55 and 1/0.7, which correspond to shock
type and cyclic type earthquake motion,
respectively. They were obtained by comparing
cyclic loading test and shear stress-time history
(Ishihara and Yasuda, 1975), and number of
effective cycles is evaluated to be less than, or
greater than or equals to three, respectively.
During the 2003 Tokachi-oki ear earthquake

a
nk
Na

Fault zone

iT

gh
rou

JMA Intensity
>6.0
5.5~6.0
5.0~5.5
4.5~5.0
<4.5

Figure 1: Fault zone of Tonankai and Nankai


earthquakes and JMA intensity

521

when these earthquake occurs simultaneously.


Seismic intensity exceeds 6 in the widespread
area, therefore liquefaction is expected in the
wide area, too. As explained later in Figure 2
(a), duration of this earthquake is very long;
much longer than that of 2003 Tokachi-oki
earthquake. Therefore, effective number of
cycles will be much larger than a few cycles that
are used in the conventional research.
Effect of different earthquake motion on
onset of liquefaction is investigated through
numerical examples.

GL-32m

Acceleration (m/s )

2
1
0
-1
-2
0

100

200

300
Time (sec.)

500

600

(a) Tonankai earthquake


GL-32m

Acceleration (m/s )

2
1
0
-1
-2
0

10

15

20

25

30
35
Time (sec.)

40

45

50

55

60

(b) Hyogoken-nambu earthquake


Figure 2. Earthquake motions at Amagasaki

EARTHQUAKE MOTION AND SITE

Ashiya city

The synthesized earthquake motion, shown


in Figure 2 (a) is used in this research, because
no earthquake record exists against the
Tonankai earthquake. This earthquake motion
is made for the site of the Research Institute of
Kansai Electric Power Cooperation located at
Amagasaki City (see Figure 3 for location). The
earthquake record was observed during the
1995 Hyogoken-nambu (Kobe) earthquake
shown in Figure 2 (b). Therefore, they are
convenient for the purpose of this research.
Unfortunately, liquefaction was not reported
during the 1995 Hyogoken-nambu earthquake,
and soil profile seems to indicate that
liquefaction will not occur. Therefore another
site must be used for comparison. The vertical
array site at Port Island, Kobe city, is employed
for this purpose. Significant soil liquefaction
occurred during the 1995 Hyogoken-nambu
earthquake; boiled sand covered the ground
widely as shown in Figure 4.
Figure 5 shows soil profiles of this site. The
earthquake motion shown in Figure 2 is applied
at GL-32 m where seismograph was installed.
Layer above GL-18m is fill made of
decomposed granite called masado. Layers
between GL-18m and GL-28m is very soft
Holocene clay called Ma13 (marine clay).
Holocene and Pleistocene gravel layer exists
below this clay layer.

Nishinomiya city
Amagasaki city

HIgashinada ward
Nada ward

Research Institute, KEPC

Chuo ward

JMA Scale 7
***
0 1 2km

Port Island

Figure 3. Map of analyzed site.

Vertical array site


Ground improvement

METHOD OF ANALYSIS

Computer code YUSAYUSA (Yoshida and


Towhata, 1991) is used to analyze this site.
Elastic modulus and nonlinear properties are
same with Yoshida (1995). Figure 6 shows
dynamic deformation characteristics and

400

Sand boil
0

500

1000

1500m

Figure 4: Sand boil during the 1995 Hyogokennambu earthquake at Port Island and vertical
array site

522

modeling by Hardin-Drnevich model (Hardin


and Drnevich, 1972).
Hyperbolic models are used as nonlinear
characteristics whose reference strain is same
with Hardin-Drnevich model shown in Figure 6,
but since Masing rule is employed to build
hysteresis loop after unloading/reloading,
damping characteristics is different from Figure
6. Simulation of liquefaction strength is shown
in Figure 7.

Depth
(m)

SPT-N value

Soil Type

10 20 30 40 50

Vs
(km/s)

Vp
(km/s)

0.17

0.26

-5

0.33

Fill
(Masado)

-10

0.21

0.78
1.48

-15
-20
-25
-30
-35

RESULT AND DISCUSSION

Holocene
clay
(Ma13)

0.18

1.18

Holocene
gravel

0.245

1.33

0.305

1.53

0.350

1.61

0.303

1.61

0.32

2.0

-40
-45

Maximum response by two earthquakes is


shown in Figure 8. PGAs are about 1.5m/s2 by
two analyses; accelerations yield upper limit
because soil failed or stress becomes nearly
shear strength. This can be seen from the
acceleration time history at the ground surface
shown in Figure 9; time history shows constant
acceleration range at several maximums.
Looking at maximum excess porewater
pressure, liquefaction does not occur under the
Hyogoken-nambu (Kobe) earthquake and it
occurs under the Tonankai earthquake.
Difference between the liquefaction during the
Kobe earthquake (liquefied) and this analysis
(non-liquefied) can be explained by considering
that the Amagasaki city where earthquake
motion used in this analysis is observed is more
far from the epicenter than Port Island resulting
in smaller earthquake motion.
The earthquake motion of Hyogoken-nambu
earthquake seems stronger than that of
Tonankai earthquake because maximum shear
stress at Holocene clay layer (Ma13) is larger
during Hyogoken-nambu earthquake than
Tonankai earthquake (see Figure 10, stress-

Pleistocene
gravel

-50
-55
-60
-65

Pleistocene
clay
(Ma12)

-70
-75
-80

Pleistocene
gravel
seismograph

Figure 5: Soil profile

1.0
Clay
0.8
0.6
0.4

20
Test
Model

10

0.2
0
0.0001

0.001

0.01

0.01

0
10

Strain, (%)

(a) Holocene clay


1.0

Shear stress ratio, d /'v0

0.4

Fill

0.8

Test
Simulated

0.6

0.3

0.4

0.2

0.2

0.1

0
0.0001

20

Disturbed
Undisturbed
Model

0.001

0.01

10

0.01

0
10

Strain, (%)

0.0
1

(b) Fill
Figure 6: Dynamic deformation characteristics

10
100
Number of cycles causing liquefaction

Figure 7: Liquefaction strength

523

overburden stress, which indicates liquefaction


does not occur. On the other hand, maximum
excess porewater ratio reaches initial
overburden stress, i.e., liquefaction occurs.
In the conventional simplified method to
evaluate onset of liquefaction (e. g. Japan Road
Association (1996) and AIJ (2001)), so called FL

strain curves at the 23rd layer, GL-28~30 m).


Maximum shear stress is larger not only in
Ma13 layer but masado (fill) layer under
Hyogoken-nambu earthquake than Tonankai
earthquake. Maximum excess porewater
pressure under Hyogoken-nambu earthquake,
however, does not reach initial effective
Max. Acceleration
t
Soil Vs
2
(m/sec )
Type (m/s) (tf/m3)
1
2
3

Depth
(m)
1.0
2.0
3.0
4.0
5.0
6.0
7.0

Fill

8.0
9.0
10.0
11.0
12.0
13.1
14.2
15.3
16.4
17.2
18.0

117.1

1.7

154.2

1.7

175.2

1.7

173.0

2.0

180.3

2.0

186.8

2.0

192.6

2.0

198.0

2.0

203.0

2.0

207.6

2.0

211.9

2.0

216.0

2.0

220.1

2.0

224.1

2.0

227.9

2.0

231.5

2.0

234.5
237.0

2.0
2.0

174.4

1.7

177.3

1.7

180.0

1.7

182.6

1.7

185.1

1.7

236.2

2.0

240.8

2.0

Max. Displacement
(cm)
5

10

Max. Stress
2
(kN/m )

15

20

40

Max. Strain
(%)
-1

60

Excess PWP
2
(kN/m )
1

10

10

200

20.0
22.0

Ma13
24.0
26.0
28.0
30.2 Gravel

Kobe eq.
Tonankai eq.

32.4

Figure 8: Maximum response

1
0
-1
-2
0

100

200

(a) Tonankai eq.

300
Time (sec.)

400

500

Acceleration (m/s )

Ground surface

Acceleration (m/s )

600

0
-1
-2
0

Shear stress, (kPa)

Shear stress, (kPa)

60

Ma13

20
0
-20
-40
-60
-0.2

GL-29.0m
-0.1

10

15

20

25

30
35
Time (sec.)

40

45

50

55

(b) Hyogoken-nambu eq.


Figure 9: Acceleration at ground surface

60
40

Ground surface

0
0.1
0.2
Shear strain, (%)

Ma13

40
20
0
-20
-40
GL-29m

-60

0.3

-0.2

0.2
0.4
Shear strain, (%)

(a) Tonankai eq.


(b) Hyogoken-nambu eq.
Figure 10: Stress-strain curves of Holocene clay layer

524

0.6

60

Development of excess porewater or decrease


of effective overburden stress stops at certain
stage because stress amplitude becomes
smaller. On the other hand, although stress
amplitudes are similar between cases of the
Hyogoken-nambu earthquake and Tonankai
earthquake, the case of the Tonankai

method, liquefaction is easier to occur if


maximum shear stress is larger. Result of the
analysis shows, however, quite opposite
tendency.
This inconsistency can be recognized from
Figure 11 where stress paths and stress-strain
curves at the 18th layer is compared.

Masado

20
0
-20
-40

20
0
-20
GL-17.6m

-40

GL-17.6m

Masado

40
Shear stress, (kPa)

Shear stress, (kPa)

40

50
100
150
200
Effective overburden stress, 'v0 (kPa)

-2

-1

0
1
2
3
Shear strain, (%)

(a) Tonankai Eq.


60

60

Masado
Shear stress, (kPa)

Shear stress, (kPa)

40
20
0
-20
-40
GL-17.6m

-60
0

Masado

40
20
0
-20
-40
GL-17.6m

-60

50
100
150
200
Effective overburden stress, 'v0 (kPa)

-0.2

0
0.2
Shear strain, (%)

0.4

(b) Hyogoken-nambu eq.


Figure 11: Stress path and stress-strain curve in 18th layer (GL-17.2~18.0 m)
1000

Kobe eq.
Tonankai eq.
2

Acceleration (cm/s )

Amplification factor

100

10

0.1
0.1

1
Frequency (Hz)

100

10
Kobe eq.
Tonankai eq.
1
0.1

10

1
Period (sec.)

(a) Amplification ratio


(b) Response spectrum
Figure 12: Spectral response

525

10

Earthquake Engineering Symposium, Tokyo, pp.


641-648

earthquake shows much larger cycles of


loading. The effective overburden stress
decrease very slowly, but constantly. After
reaching the phase transform line, effective
overburden stress decreases quickly, resulting
in liquefaction. It is, therefore, obvious that large
number of cycles in the Tonankai earthquake is
the main reason why liquefaction occurred.
Figure 12 compares amplification ratios and
acceleration response spectra (5% damping) by
two earthquakes. Amplification is larger in the
Tonankai earthquake in almost all frequency
region. On the other hand, response
acceleration is larger in Kobe earthquake in
period larger than 0.8 seconds. This indicates
that liquefaction reduces total power of the
earthquake motion at the ground surface.

Japan Road Association (1996): Specifications for


Highway Bridges, Part V, Seismic design (in
Japanese)
Yoshida, N. and Towhata, I. (1991): YUSAYUSA-2
and SIMMDL-2, theory and practice, revised in
2003 (version 2.1), Tohoku Gakuin University
and
University
of
Tokyo;
http://boh0709.ld.infoseek.co.jp/
Yoshida, N. (1995): Earthquake response analysis at
Port Island during the 1995 Hyogoken-nanbu
earthquake, Tsuchi-to-Kiso, Vol. 43, No. 10, pp.
49-54 (in Japanese)

CONCLUDING REMARKS

Effect of long duration earthquake such as


ocean trench earthquake on onset of
liquefaction is investigated by comparing the
response between near field earthquake
(Hyogoken-nambu earthquake) and coming
ocean trench type huge earthquake (Tonankai
earthquake). Since ocean trench type
earthquake have much larger number of cycles,
liquefaction is easier to occur. This effect is not
taken into account in the conventional design
specifications.
REFERENCES
AIJ (2001): Recommendations for design of building
foundations, 2001 Revision
Central Disaster Prevention Council, Cabinet Office
of Japan (2005): Document at the 6th meeting of
expert committee on Tonankai and Tokai
earthquakes, et al.
Hardin, B. O. and Drnevich, V. P. (1972): Shear
modulus and damping in soils: design equations
and curves, Proc. of the American Society of civil
engineers, Vol. 98, No. SM7, pp. 667-692
Ishihara, K. and Yasuda, S. (1975): Effect of
irregularity of earthquake motion and initial
confining stress on liquefaction, Tsuchi-to-Kiso,
Vol. 23, No. 6, pp. 26-35
Iwasaki, T., Tatsuoka, F., Tokida, K. and Yasuda, S.
(1978): A practical method for assessing soil
liquefaction potential based on case studies at
various sites in Japan, Proc., 5th Japan

526

A simple evaluation method for earthquake damage to the quay walls


M. Soejima, A. Suizu, J. Ejiri, T. Matsuda
Technical Research Institute of OBAYASHI Co., Japan

Abstract
ABSTRACT: In Japan, a lot of reclaimed lands had been constructed for port and industry
facilities since the head of the 20th century. A reclaimed land is generally composed of both
the quay walls and the backyard ground. A key of the seismic safety of reclaimed lands is
sustainability of the quay walls. A simple evaluation method for earthquake damage to the
quay walls has been constructed empirically based on actual earthquake damage by such
the past earthquakes as 1995 Hyogo-ken Nanbu and other moderate magnitude
earthquakes. This method is limited to be applied in order to pick up the weak quay walls
with less seismic performance.
INTRODUCTION

A lot of the quay walls which locate


alongshore on reclaimed land having the
important port and industry facilities have not
sufficient seismic performance, because they
were constructed in the old age.
During 1995 Hyogo-ken Nanbu earthquake,
many gravity quay walls suffered heavy
damage as shown in Fig 1 and Fig 2.
Representative damage to gravity quay walls
involves large seaward displacement and
settlement with the value of several meters
accompanying large settlement of backyard
grounds caused by liquefaction of backfill sand.
It is known that the shaking even with the
JMA seismic intensity scale did damage to
the quay walls in the past earthquake.
When the port facilities like the quay walls
suffer heavy damage and the access route to
damaged area from the sea is cut off, such
restoration resources as water, food, medical
materials and equipments can not be
conveyed smoothly just after earthquake
attack. So, it is quite important thing that
sufficient seismic performance is guaranteed
to the quay walls. In addition, it is pointed out
that a simple evaluation method of earthquake
damage to the quay walls is necessary in
order to pick up efficiently the weak quay walls
with less seismic performance, because
several types of the quay walls spread
alongshore with long distance and cross the
various ground conditions.

527

In this paper, a simple evaluation method of


earthquake damage to the quay walls has been
proposed based on actual earthquake damage
by such the past earthquakes as 1995 Hyogoken Nanbu earthquake and other recent
moderate magnitude earthquakes.

Fig 1: Large seaward displacement of quay wall

Fig 2: Large settlement of backyard ground

A SIMPLE EVALUATION METHOD

Fig 3 shows the flow of evaluation of


earthquake damage to the quay walls by this
method considering the damage level as
horizontal seaward displacement D at the top of
the quay walls as shown in Fig 4. Major
parameters used for evaluation are quay walls
structural types, peak horizontal ground velocity
PGVs on backyard ground surface, thickness of
liquefied backfill sand in backyard ground,
liquefaction potentiality of the substitution sand
mat beneath the quay walls. Two structural
types are considered, one is gravity type and
another is steel sheet pile type.
If all parameters are same condition except
structure types, evaluated damage of sheet pile
type is slightly heavier than gravity type.
For determination of PGVs, non-linear one
dimensional site response analysis is used to
be conducted referring the boring log data of
backyard ground.
Liquefaction analysis of Specification for
Highway Bridge (2002) is adopted to estimate the
thickness of liquefied backfill sand using vertical
distribution of shear stress by above-mentioned
non-linear one dimensional site response
analysis. The thickness of liquefied backfill sand
is defined as total thickness of sub layers with
FL value of less than 1.0.
Structural type
Gravity system
Sheet-pile system

If we can not get the physical properties of


the substitution sand on site, those of the Port
island and the Rokko island are used where the
substitution sand was experienced liquefaction
during 1995 Hyogo-ken Nanbu Earthquake, with N
value from 5 up to 15.
Table 1 shows the damage rank by D on
the point of restoration.
Horizontal seaward
displacement D
Backyard ground

H
Sea side

A part
Max.
20m

B part (20H)
Fig 4: Definition of horizontal seaward displacement
D
Table 1: Damage rank of the quay walls
D(cm)
0
0 - 25
25-70
70-200
more than 200

Backyard ground condition


Ground categories
Vs ,N-value ,
G-,h-

Damage rank
0

Damage state
No damage
A little repair
Considerably damaged
Heavily damaged
Collapse

Input Ground Motion on


engineering-oriented base

Non-linear one dimensional site response analysis


Vertical distribution of shear stress in backyard ground
Ground-state index for
each quay wall types
(Table 4 and 5)

Liquefaction analysis
Vertical distribution of Fl values
Liquefied conversion thickness

Peak horizontal
velocity PGVs on
backyard ground
surface

Calculation of deformation rates (Table2 and 3)


Horizontal seaward displacement D at the top of the quay walls
D=1.21H/100
Damage rank (Table1)

Fig 3: the flow of evaluation for earthquake damage to the quay walls

528

The following equations are used to calculate


horizontal seaward displacement D (cm) and
deformation rate (%).

0 is determined from the relationship


between ground-state index and peak
horizontal ground velocity PGVs (cm/s) on
backyard ground surface shown in Table 3.
Table 4 and 5 show the ground-state index
for gravity type and sheet pile type respectively
which represent the relationship between
liquefaction potentiality and its areas of
backyard ground. C Part in Table 5 shows
anchor of tie rod of sheet pile type. Ground
condition of C part is considered too.
Above all parameters are calibrated and
constrained to actual earthquake damage data
from such the past earthquakes in Japan as
1968 Tokachi-oki,
1973 Nemuro-hanto-oki,
1983 Nihonkai-chubu, 1993 Kusiro-oki, 1993
Hokkaido-Nansei-oki, 1995 Hyogo-ken Nanbu.

D=1.21H /100
Here,
1: Final deformation rate (%),
H : Quay wall height (cm),
0: Initial deformation rate (%).
The factor of 1.2 expresses the safety factor
covering uncertainty of ground information.
Observed deformation rate from actual
damage is defined as D/H100 (%).
Table 2 represents the relationship between
deformation rate 1 and 0 for each groundstate index.
Table 2: Final deformation rate
1
Ground-state index

Table 4: Ground-state index for gravity type

Final deformation rate 1 (%)

Total thickness of FL1.0 sublayers

B: Thickness of liquefied total soil layer


B 2m 1=0.50
2mB5m 1=0.80
5mB
1=1.50

G 1,G, S1
S 4, S
G, G2, S1
S2, S2, S3

Less than1/2H'

More equal to 1/2


H'

G2

G1

FL 1.0
(Non liquefaction)
FL 1.0
(Liquefaction)

1= 0

Table 3: Initial deformation rate 0


Ground-state index

PGVs10

10PGVs25

25PGVs50

50PGVs100

100PGVs

0.1PGVs

10

G1G2

0.2PGVs

20

0.4PGVs

40

S1S2

0.2PGVs -5

15

S1S2

0.2PGVs -2

0.48PGVs -9

0.1PGVs +10

20

S3S4

0.4PGVs -4

0.96PGVs -18

0.2PGVs +20

40

PGVs -10

1.4PGVs -20

0.5PGVs +25

75

Table 5: Ground-state index for sheet pile type


Total thickness of FL 1.0 sublayers
Less than 1/2H'
More equal to 1/2H'
FL1.0

There is the part


of FL1.0
FL1.0
(Thickness less
than 2m)
FL1.0
(Thickness more
equal to 2m)

Rigid anchor structure and


Non liquefaction in C part

S1

S2

Weak anchor structure and


Liquefaction in C part

S1

S3

Rigid anc hor structure and


Non liquefaction in C part

S2

S4

S1

S3

S1

Weak anchor structure and


Liquefaction in C part

529

(cm)

VERIFICATIONS

Gravity system
Sheet pile system

Evaluated

Evaluated D

Fig 4 shows comparison of D values


between Observed values of previous 5
earthquakes except 1995 Hyogo-ken Nanbu
earthquake and Evaluated values by proposed
evaluation method. D values were caused by
the maximum excitation of the JMA seismic
intensity scale . D scatters with small values,
however Evaluated D values are proportional to
Observed D values.
Fig 5 presents comparison of values
(cm)
between Observed one of 1995 Hyogo-ken
Nanbu earthquake and Evaluated one.
values were caused by the maximum
excitation of the JMA seismic intensity scale
Observed D of 5 earthquakes except 1995
,where Observed D values ranged from 1.5 to
6.0 meters. Evaluated values have good Fig 4: Comparison between Observed and Evaluated D
agreement with Observed D values.
It has been assured that proposed a simple
(%)
evaluation method in this study has appropriate
Gravity system
accuracy for picking up the weak quay walls with
less seismic performance.
CONCLUSION

A simple evaluation method for earthquake


damage to the quay walls has been proposed
here based on actual damage data from 1968
Tokachi-oki, 1973 Nemuro-hanto-oki, 1983
Nihonkai-chubu,
1993 Kusiro-oki,
1993
Hokkaido-Nansei-oki, 1995 Hyogo-ken Nanbu
earthquakes in Japan
Evaluated D or values by this proposed
simple evaluation method show good
accordance with Observed D or values.
This proposed simple evaluation method is
very useful for picking up the weak quay walls
with less seismic performance effectively.
After screening the weak quay walls with less
seismic performance, immediately, we can
discuss the efficient counter measures through
the earthquake response analysis and the largescale shaking table tests and the centrifuge test,
based on structural types of the quay walls and
precise ground conditions and seismic intensity on
specific site.
Most important things are to accumulate
the actual reliable damage data in the future
earthquakes, to keep doing the verification and
improving a evaluation method by them.

(%)
Observed of 1995 Hyogo-ken Nanbu
Fig 5: Comparison between Observed and Evaluated
REFERENCES
Iai, S., et al. (2001) Seismic Design Guidelines for
Port Structures, International Navigation
Association, ISBN 90 165 1818 8, pp.7-126
Inatomi, T., et al. (1995) Mechanism of Damage to
Port Facilities during 1995 Hyogo-ken Nanbu
earthquake (in Japanese), TECHNICAL NOTE
OF THE PORT AND HARBOUR RESEARCH
INSUTITUTE, MINISTRY OF TRANSPORT, Vol.
813, pp.51-166
Japan Road Association (2002) Specification for
Highway Bridge, Part V. Earthquake Resistant
design (in Japanese), pp.119-12

530

Effect on records of seismic intensity-meter by adjacent building


Shunichi Kataoka1
Hirosaki University, Japan

Abstract
The seismic intensity obtained from Aomori seismic intensity information network (ASIIN) is
compared with closely located K-NET. ASIIN stations are installed around buildings, while
condition of K-NET station is thought to be free-field. We first compare the difference of seismic
intensity by ASIIN and K-NET, then calculate spectral ratio between them. The absolute value
of the difference is smaller than 0.3. Amplitude of the spectral ratio is almost unity from 0 Hz to
4Hz without depending on the site. There is a peak around 5 Hz in some sites. In frequency
rage of higher than 5 Hz, spectral ratio decays gradually. We recorded an microtremor at a site
and reveal that peak in spectral ratio is induced by adjacent building vibration at that site.

preferable. Prefectural government considered


this fact then installed seismic intensity-meter
near the building.
In this paper, the seismic intensity obtained
from Aomori seismic intensity information
network (ASIIN) is compared with K-NET data
to evaluate the effect of adjacent building.
Aomori is a prefectural name of Japan. Aomori
Prefecture is located at northern tip of Honshu,
that is Main island of Japan.

INTRODUCTION

After an occurrence of a felt earthquake,


Japan
Meteorological
Agency
(JMA)
immediately issues seismic intensity information
based on the data observed by seismic
intensity-meter (Nishimae 2004). The JMA
seismic intensity is very familiar index of
earthquake severity for citizen in Japan. Public
sectors start their countermeasures for
earthquake disaster by seismic intensity value.
In Japan, there are about 3400 seismic
intensity observation sites. JMA deploys about
600 sites and local prefectural governments
have about 2800 sites. In each prefecture,
almost seismic intensity-meters are installed
near municipal government offices. Ordinary, a
separation is small, an effect of adjacent
building on the seismic intensity-meter is
anxious. On the contrary, some K-NET,
Kyoshin-network (Aoi et al, 2004) station are
placed in premises of municipal governments
office. In those case, K-NET is installed end of
the premise and there is no buildings around KNET. This situation is like a free-field condition.
Strictly to say, at first K-NET stations have
been installed at those sites and seismic
intensity calculated from K-NET had been used.
However, old type K-NET (K-NET95) could not
transmit seismic intensity during an earthquake
but after an earthquake. As seismic intensity is
very
crucial
information
for
mitigating
earthquake disaster, quick response is more

INSTRUMENTAL SEISMIC INTENSITY IN JAPAN

In Japan, seismic intensity is determined by


a instrument. The process from threecomponent ground acceleration to seismic
intensity is outlined in Fig.1. At first, threecomponent ground acceleration is recorded.
Then Fourier transform is performed to use a
filter. The Filter characteristics is displayed in
Fig.2. The peak frequency of the filter is 0.5Hz
and amplitude of the filter becomes small
gradually as frequency becomes high. After
filtering, inverse Fourier transform is performed
to obtain time series. Sum of squares of filtered
three component ground motion is used to
calculate the instrumental seismic intensity
scale. A threshold amplitude (a0) that lasts for
0.3 seconds is substituted in equation (1).
Finally, we can get the instrumental seismic
intensity scale.
I=2 log (a0)+0.94

531

(1)

Acceleration
N-S comp.

Acceleration
E-W comp.

Acceleration
U-D comp.

Fourier
transform

Fourier
transform

Fourier
transform

Filtering

Filtering

Filtering

IFT

IFT

IFT

table 1. There is no building around the station


and ground surface is flat. ASIIN stations are
also investigated. At Towada and Kawauchi,
sensor is installed inside of the office building
and for the other sites, sensor is installed out
side of office building with separation of about
one or two meters.
Earthquakes used in this study are listed in
Table 2. Epicenters and observation sites are
displayed in Fig. 3. As Fig. 3 shows that the
epicentral distance is very large comparing to
the distance between K-NET and ASIIN, we can
compare the ground motion obtained at each
station directly.

Sum of squares
Determine the peak value

DIFFERENCE OF SEISMIC INTENSITY SCALE

Modified Kawasumi's formula

Instrumental seismic intensity scales are


calculated from ground motions recorded by KNET. Then, compare to those of ASIIN
published data at each site. Fig. 4 shows
examples of comparing. The horizontal axis
corresponds to seismic intensity comes from KNET and the vertical axis corresponds to ASIIN
seismic intensity. The left plate of Fig. 4 is for
the case of Kawauchi. In this case, an average
of the difference is -0.1 and standard deviation
is very small. This bias may corresponds to

Instrumental seismic
intensity scale

Fig 1: Outline of obtaining instrumental seismic


intensity scale.

Fig 2: Characteristics of the filter to obtain


instrumental seismic intensity.

where I is instrumental seismic intensity.


STUDY SITES AND EARTHQUAKES

Study sites are listed in table 1. At those


sites,there are K-NET and ASIIN station very
closely and those geomorphologic units
(Wakamatsu et al , 2005) are the same.
The author visited to check environmental
condition around all K-NET stations listed in

Fig 3: Map showing observation sites (triangle) and


epicenters (star).

532

Table 1: Study sites.


Site name

K-NET
Code

Elevation [m]
K-NET

Separation
distance [m]

ASIIN

Geomorphologic unit

Imabetsu

AOM025

17

Nishimeya

AOM029

120

120

90 Gravelly terrace

Ikarigaseki

AOM015

144

144

70

Towada

AOM021

68

72

260

Gravelly terrace

Misawa

AOM011

41

42

390

Terrace covered with volcanic ash

Noheji

AOM010

15

23

1500

Terrace covered with volcanic ash

Nanbu

AOM013

30

30

140

Kawauchi

AOM006

70

250

Gravelly terrace
Valley bottom lowland

Hill
Marine sand and gravel bars

Table 2: Earthquake parameters.


No.

Date

EQ.1
EQ.2
EQ.3
EQ.4
EQ.5
EQ.6
EQ.7
EQ.8
EQ.9

2002/02/14
2002/04/04
2002/08/12
2002/10/14
2002/12/01
2003/04/17
2003/05/26
2003/09/26
2003/09/26

HH:MM
10:12
8:42
6:55
23:12
18:57
2:59
18:24
4:50
6:08

Lat.[]

Long.[]

41.4633
41.4700
41.0833
41.1483
42.6600
40.9567
38.8050
41.7783
41.7067

142.0650
142.0117
142.6383
142.2800
143.9683
142.3450
141.6817
144.0783
143.6950

Depth
[km]
64
59
26
53
103
40
71
42
21

M
5.0
5.3
5.0
6.1
5.4
5.6
7.1
8.0
7.1

Fig 4: Instrumental seismic intensity scales obtained by K-NET and ASIIN.

some physical phenomena. Effect of building is


the highest possibility. However, for the case of
right hand plate of Fig. 4, though mean is
almost equal to zero, the variance is large.
These comparisons are summarized in
Table 3 as the reference is K-NET. Mean

values are scattered from -0.30 to +0.15. More


than half sites have negative mean value.
However, seismic intensity is broadcasted as
round-off value to the public, these differences
are not significant for disaster information. From
this study, we can conclude that closely located

533

Table 3: Summary of seismic intensity differences.


Difference in
Spectral study
Number
seismic intensity
Site name
of EQ
Mean
S.D.
EQ7
EQ8
EQ9
Imabetsu
4
-0.03
0.05
Nishimeya
4
0.15
0.10
OK
OK
Ikarigaseki
4
-0.13
0.05 OK
OK
OK
Towada
7
0.10
0.10
Misawa
8
-0.30
0.14
OK
OK
Noheji
8
0.01
0.20
Nanbu
8
-0.26
0.17 OK
OK
OK
Kawauchi
8
-0.10
0.05
OK
OK

sites with the same geological condition indicate


almost the same instrumental seismic intensity.

there is one trough at 7Hz.


As spectral ratio does not depend on
events, an average of ratio is calculated for
every sites then displayed in Fig. 6. We can
recognize that spectral ratios of each site are
flat below 4Hz.
Seismic intensity difference can be
explained by
spectral ratio shape. At
Kawauchi, the ratio is almost equals to unity,
and absolute value of the difference is the most
minimum as 0.10 among five sites. Negative
difference may comes from a trough around
4Hz. At Kawauchi, the sensor of ASIIN is
installed in a building. However, we can not find
out a effect of the building on the spectral ratio.
The building is made of wood and lighter than
ordinal RC structure. A light weight of the
building is one of the reasons.
At Ikarigaseki, spectral ratio has a peak
around 5Hz, while there are some troughs.
Peak and troughs influence small amount of the
difference.
For Nanbu, it has peak and big trough. For
Misawa, there is no peak and amplitude decays
in higher frequencies. These tendency

SPECTRAL RATIO

To study the difference of seismic intensity


more precisely, earthquake ground motion must
be used. The intensity-meter of ASIIN can store
the ground motion. However, data in the
instrument
is
not
collected
routinely.
Additionally, data is overwritten in the case of
memory full. So it is not easy to deal ground
motion data. The result of collecting data is also
displayed in Table 3. We can collect only EQ7,
EQ8, and EQ9's ground motions at five sites.
Spectral ratios of ASIIN to K-NET at
Ikagiraseki site are displayed in Fig. 5. At this
site we can use three event record. Spectral
ratios of each event are almost the same in
both NS component and EW component. In NS
component, ratio is almost unity below 4Hz. At
around 5Hz, there is a peak and there are two
troughs at about 5.5Hz and 7Hz. In EW
component, shape is almost the same to the
one of NS component below 5Hz. However,

Fig 5: Spectral ratio of ASIIN to NET at Ikarigaseki site.

534

Fig 6: Averaged spectral ratio of ASIIN to K-NET at all sites.

correspond to rather large differences of the


intensity.
At Nishimeya, spectral ratio is larger than
unity at least in the frequency range of 0 to 7Hz
and there is small trough. That is a reason why
the difference is positive. A Quantitative study
of spectral ratio and difference of seismic
intensity is performed by Kataoka et al(2004).
MICROTREMOR STUDY

At Ikarigaseki, Nanbu, and Nishimeya, there


is a peak around 5Hz. To clarify a origin of the
peak, microtremor test is performed at
Ikarigaseki. Ikarigaseki municipal government
office building has three story and no basement
floor. Microtremor is recorded at top and ground
level simultaneously and calculated spectral
ratio between them. Fig. 7 shows the spectral
ratios. Both in NS and EW component, there is
a peak at 5Hz. So that peak in Fig. 5 and Fig. 6
corresponds to building induced vibration. In
another word, this is the result of the inertial
soil-structure interaction.
Microtremor at free-field and on the
container of ASIIN is also recorded
simultaneously. The spectral ratios are
displayed with those of earthquake motion in

Frequency [Hz]
Fig 7: Amplification of the building at Ikarigaseki
site estimated from microtremor.

Fig. 7. Shapes of spectral ratios are very similar


to each other. An amplitude of spectral ratio is
smaller than unity in higher frequency. We
conclude that decrease of amplitude of spectral
ratio is due to the kinematic soil-structure
interaction.

535

Earthquake
Microtremor

Earthquake
Microtremor

(a) NS comp
Frequency [Hz]

Earthquake
Microtremor

(c) UD comp

(b) EW comp
Frequency [Hz]

Frequency [Hz]

Fig 8: Comparing spectral ratio of ASIIN to K-NET ( free-filed) calculated from earthquake motion and
microtremor.
CONCLUDING REMARKS

vibration. Decreasing amplitude of spectral ratio


is derived from the kinematic soil-structure
interaction.

In Japan, seismic intensity is very important


and familiar index. However, circumstance
around the seismic-intensity meter has not been
considered seriously. Some seismic-intensity
meters operated by local government are
installed around buildings. In this paper,
building affection on the seismic intensity is
studied used observational value with
comparing K-NET data. We use seismic
intensity obtained by Aomori seismic intensity
information network (ASIIN) as a case study. KNET data is a representation of free-filed
ground motion. Main results obtained in this
study are summarized as follows.
Closely located sites with the same
geological condition indicate almost the same
instrumental seismic intensity. The difference
scattered from -0.30 to +0.15.
Seismic intensity difference can be
explained by spectral ratio of ASIIN to K-NET.
In this paper, only 5 sites are studied with 2 or 3
earthquake ground motions. However, spectral
ratios are stable. Amplitude of the spectral ratio
is almost unity from 0 Hz to 4Hz without
depending on the site. There is a peak around 5
Hz in some sites. In frequency rage of higher
than 5 Hz, spectral ratio decays gradually.
Peaks and troughs control the difference of
seismic intensity.
From microtremor test at Ikarigaseki site,
peak in spectral ratio is due to buildings

REFERENCES

Aoi S., Kunugi T., and Fujiwara H. (2004)


Strong-motion seismograph network
operated by NIED: K-NET and KiK-net,
Journal of Japan Association of Earthquake
Engineering, vol.4, No.3, 65-74.
Kataoka, S., Ichimura T., and Kikuchi, T. (2005)
Characteristics of the seismic intensity
obtained from the Aomori seismic intensity
information network by comparing to the KNET data, JSCE Journal of Earthquake
Engineering, vol.28, C00086.pdf (in
Japanese with English abstract)
Nishimae Y. (2004) Observation of seismic
intensity and strong ground motion by Japan
meteorological agency and local
governments in Japan, Journal of Japan
Association of Earthquake Engineering,
vol.4, No.3, 75-78.
Wakamatsu, K., Kubo, S., Matsuoka, M.,
Hasegawa, K., and Sugiura, M (2005) Japan
Engineering Geomorphologic classification
Map, University of Tokyo Press.

536

Anda mungkin juga menyukai