You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/268590571

Seismically Induced Lateral Earth Pressures on Retaining Structures and


Basement Walls

Conference Paper  in  Geotechnical Special Publication · May 2012


DOI: 10.1061/9780784412138.0013

CITATIONS READS
24 2,288

3 authors:

Nicholas Sitar Roozbeh Geraili Mikola


University of California, Berkeley WSP USA
178 PUBLICATIONS   5,011 CITATIONS    28 PUBLICATIONS   225 CITATIONS   

SEE PROFILE SEE PROFILE

Gabriel Candia
Universidad del Desarrollo, Santiago, Chile
31 PUBLICATIONS   105 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Ocean Beach Master Plan: Coastal Management Framework View project

Evaluación del Riesgo Hidrometeorologico de Cultivos en la Region de la Araucania View project

All content following this page was uploaded by Gabriel Candia on 27 September 2016.

The user has requested enhancement of the downloaded file.


Seismically Induced Lateral Earth Pressures on Retaining Structures and
Basement Walls

Nicholas Sitar1, Roozbeh Geraili Mikola2, and Gabriel Candia3


1
Edward G. Cahill and John R. Cahill Professor, Dept. of Civil and Environmental
Engineering, UC Berkeley, Berkeley, CA 94720-1710, email: sitar@ce.berkeley.edu
2
Ph.D. candidate, Dept. of Civil and Environmental Engineering, UC Berkeley,
Berkeley, CA 94720-1710, e-mail: roozbehg@berkeley.edu
3
Ph.D. candidate, Dept. of Civil and Environmental Engineering, UC Berkeley,
Berkeley, CA 94720-1710, e-mail: gacandia@berkeley.edu

ABSTRACT

The evolution of the methodology for the evaluation of seismic earth pressures on
retaining structures and basement walls is reviewed together with observations of
field performance. The case history data and data from recent experimental work are
used to show that the currently used methods are quite conservative and lead to
excessively conservative designs in regions where design PGA exceeds 0.4g.
Specifically, the experimental data from seismic centrifuge tests shows that the
seismic earth pressure distribution for moderate size retaining structures, on the order
of 6-7 m high, is triangular, increasing with depth. Moreover, there is no significant
increase in seismic earth pressure between unbraced and braced structures with fixed
base, while the loads on free standing cantilever structures are substantially lower
owing to their ability to translate and rotate. The significance of the observed seismic
earth pressure distributions is that the dynamic force can be applied at 1/3H, as is
done for static loading, which substantially decreases the design level seismic
moments on the structures.

INTRODUCTION

The problem of evaluating seismically induced lateral earth pressures on retaining


structures has been first addressed in the 1920’s in pioneering research carried out in
Japan by Okabe (1926) and Mononobe and Matsuo (1929). Since then this problem
has received periodic attention from the research community (e.g. Seed and Whitman,
1970; Nazarian and Hadjian, 1979; Prakash et al., 1969; Prakash, 1981; and Aitken,
1982); however, it had relatively little impact on design and engineering practice until
relatively recently.
In the United States the Uniform Building Code (UBC) did not contain
provisions for seismic design of retaining structures until 2003, although the
California Building Code (CBC) contained provisions for certain types of building
walls going back to 1980’s (Lew et al., 2010b). Since then, however, the various
provisions and recommendations have become more explicit and stringent, especially
as it comes to the recommended method of analysis and the estimation of the design
accelerations. The first comprehensive document to address this issue is the FEMA

1
450 document: “NEHRP Recommended Provisions for Seismic Regulations for New
Buildings and Other Structures” (NEHRP 2004) which has now been updated as
FEMA 750 (NEHRP 2010). Both documents endorse the use of the M-O solution or
the M-O solution as simplified by Seed and Whitman (1970) for “yielding walls” and
the Wood (1973) solution for “non-yielding” walls. However, while Seed and
Whitman (1970) use 0.85 PGA as the effective acceleration and it has been a general
practice to use 0.67 PGA for design, the FEMA 750 document states:
“In the past, it was common practice for geotechnical engineers to reduce the
instantaneous peak by a factor from 0.5 to 0.7 to represent an average seismic
coefficient for determining the seismic earth pressure on a wall. The reduction
factor was introduced in a manner similar to the method used in a simplified
liquefaction analyses to convert a random acceleration record to an
equivalent average series of cyclic loads. This approach can result in
confusion on the magnitude of the seismic active earth pressure and,
therefore, is not recommended. Any further reduction to represent average
rather than instantaneous peak loads is a structural decision and must be an
informed decision made by the structural designer.”
The immediate impact of this recommendation is on the applicability of the
M-O method, since the earth pressure computed with the M-O method increases
exponentially leading to very large forces at accelerations in excess of 0.6 g as
discussed later. The practical consequence is that the M-O method cannot be directly
applied in areas of high seismic demand and in practice requires a “work around”,
usually the addition of a small amount of cohesion. The Seed and Whitman (1970)
approximation does not suffer from the same shortcoming as it is a straight line
approximation to the M-O solution. On the other hand, this method recommends that
the seismic earth pressure increment be applied at 0.6H, which, when combined with
the recommended direct use of PGA, results in very large computed moments at the
base of the retaining structure.
At this point the questions that have to be raised are: “What evidence do we
have that indeed there is a problem requiring such step up in design
recommendations?” and “Are the analysis and design methods appropriate for the
proposed use?” The purpose of this paper is threefold: 1) address these questions
using observations of seismic behavior of retaining structures and basement walls in
past earthquakes; 2) review the theoretical and experimental basis behind the basic
limit equilibrium design methods currently used in practice; and 3) present the results
of recent experimental studies that introduce new data and suggest possible
improvements to the currently used approaches.

OBSERVATIONS FROM EARTHQUAKES

A review of the performance of basement walls in past earthquakes by Lew et al.


(2010a) shows that failures of basement or deep excavation walls in earthquakes are
rare even if the structures were not explicitly designed for earthquake loading.
Failures of retaining structures are most commonly confined to waterfront structures
retaining saturated backfill with liquefaction being the critical factor in the failures.
Failures of other types of retaining structures are relatively rare (e.g. Whitman, 1991;

2
Al-Atik and Sitar, 2010) and usually involve a more complex set of conditions,
typically, but not exclusively, involving sloping ground either above or below the
retaining structure if not both. With these general observations in mind, the purpose
herein is to illustrate the performance of retaining structures in different settings.
The photograph in Figure 1 shows failures of gravity retaining structures for a
cut slope and a railroad embankment caused by the magnitude 7.9 (7.9 USGS, up to
8.2 in other sources) Great Kanto earthquake which devastated the greater Tokyo and
Yokohama region of Japan in 1923. The retaining wall above the rail line has a slope
above it while the base of the railway embankment is in a waterway. Thus, any
number of possible failure modes can be postulated and the photograph is a good
illustration of the shortcomings of attempting to evaluate the cause and mode of
failure from photographic evidence alone. Nevertheless, it does not appear that
seismically induced earth pressure would have been the principal cause of failure of
both retaining structures, as bearing capacity and simple slope failure would also be
good candidates.

Figure 1. Failed retaining walls caused by the Great Kanto earthquake (courtesy
of the Sykes Kanto Collection, NISEE, UC Berkeley).

A more recent failure shown in Figure 2 is a gravity retaining structure along a


road cut in central Taiwan which partially failed in the 1999 Chi-Chi earthquake. This
retaining wall was a classic trapezoid shaped gravity retaining structure with no
apparent provision for seismic loading (i.e. broader section at the base). Most
importantly, however, the steep slope above the wall failed and, while that was a
seismically induced slope failure, the seismic earth pressure induced by ground
motions alone was not the cause of failure. Failures of the kind shown in Figures 1
and 2 were also observed in the 1995 Hyogoken-Nambu earthquake in Kobe, Japan
(Japanese Geotechnical Society, 1996) and they typically involved older structures
supporting poorly compacted embankments and/or on sloping ground.

3
Overall, however, there is no evidence of a systemic problem with traditional
static retaining wall design even under quite severe loading conditions (see e.g.
Gazetas et al., 2004). No significant damage or failures of retaining structures
occurred in the 1998 Wenchuan earthquake in China, or in the recent great subduction
zone generated earthquakes in Chile (2010) and Japan (2011). Figure 3 is a
photograph of a cobblestone gravity wall supporting an unfinished overpass in
Mianzhu City, China, and Figure 4 which shows a series of highway underpass
structures south of Concepcion, Chile, neither of which experienced any distress. This
observation is consistent with the conclusion reached by Seed and Whitman (1970)
who noted that gravity retaining structures designed for adequate factor of safety
under static loading should perform well under seismic loading for PGA up to about
0.3 g.

Figure 2. Trapezoidal concrete gravity retaining wall which rotated outward due
to failure of the slope above during the 1999 Chi-Chi earthquake in Taiwan (N.
Sitar photo).

Figure 3. Cobblestone gravity retaining wall for an overpass, Mianzhu City,


China, 2008 (N. Sitar photo)

4
Figure 4. Reinforced concrete cantilever walls south of Concepcion, Chile, 2010
(N. Sitar photo).
Finally, the most challenging aspect of documenting and interpreting field
performance is the fact that well documented case histories with actual design and
performance data for modern retaining structures are very sparse. A rare, well
documented case history of the performance of flood channel walls in the Los
Angeles basin during the 1971 San Fernando earthquake is presented by Clough and
Fragaszy (1977) who show that reinforced concrete cantilever structures, well
designed and detailed for static loading, performed without any sign of distress at
accelerations up to about 0.4 g, which is consistent with the previously mentioned
conclusion by Seed and Whitman (1970).

PRIOR EXPERIMENTAL DATA

Prakash (1981) identified three questions that need to be answered when


designing a retaining wall for seismic loads:
 What is the magnitude of total (seismic plus dynamic) earth pressure on the
wall?
 Where is the point of application of the resultant?
 How much has the structure been displaced?

To answer these questions numerous experimental studies have been performed in


the past and their results are very much a function of their experimental design. In
order to be able to approach the problem in a general and cost effective manner, most
of the physical model studies used either the shaking table or the geotechnical
centrifuge to model this complex, dynamic, soil-structure interaction problem.

Shaking Table Experiments. The seminal experiments using a shaking table were
performed by Mononobe and Matsuo (1929). Their original shaking table design
consisted of a rigid base box mounted on rails and driven with an ingenious conical
drum winch connected through a crankshaft to the base of the box (Figure 5). This
arrangement allowed for simple application of sinusoidal excitation with linearly
varying frequency, i.e. a frequency sweep. The ends of the box were trap doors,

5
spring mounted at the base, with pressure gauges mounted at the top to measure the
load as the “wall” tilted outward. As shown in the figure, the box dimensions were 9
ft long, 4 ft wide and 4 ft deep, with one door, door A, spanning the whole width of
the box and the other door, door B, spanning only one half of the width of the box.
Although, the box was quite substantial in size, the depth of the medium dense sand
fill was only 4 ft and the sides of the box were rigid.

Figure 5. Shaking table arrangement used by Mononobe and Matsuo (1929)

Since then numerous results of 1-g shaking table experiments have been reported
in the literature, including Matsuo (1941), Ishii et al. (1960), Matsuo and Ohara
(1960), Sherif et al. (1982), Bolton and Steedman (1982), Sherif and Fang (1984), and
Ishibashi and Fang (1987). In general, the results of these studies were in agreement
with the original results obtained by Mononobe and Okabe (1929) as far the total
resultant thrust, but a consensus emerged that the point of application of the resultant
thrust should be higher than one third the height of the wall above its base. However,
as pointed out by a number of investigators (see e.g. Ortiz et al. 1983, Stadler, 1996),
the 1-g shaking table experiments share a number of important limitations, namely:
scaling of stress and stiffness using granular, cohesionless soil is problematic under 1-
g conditions; the boundaries cannot be moved sufficiently far away from the structure
as the size of the model increases; and, the most problematic of all is the boundary at
the base of the model, since a model mounted directly on a shaking table is essentially
founded on the outcrop generating the input motion, i.e. bedrock. Hence, a retaining
structure founded on a compliant foundation cannot be readily modeled on a shaking
table at 1-g due to scaling limitations.
Centrifuge Model Experiments. Centrifuge models have the advantage that they
avoid the obvious limitations of the 1-g shaking table and scaling laws can be
accurately followed. However, they also present challenges particularly when it
comes to the ability to of conventional instrumentation to adequately respond at
frequencies dictated by the scaled model response characteristics, as discussed later.

6
Dynamic centrifuge tests on model retaining walls with dry and saturated
cohesionless backfills have been performed by Ortiz (1983), Bolton and Steedman
(1985), Zeng (1990), Steedman and Zeng (1991), Stadler (1996), and Dewoolkar et
al. (2001). Bolton and Steedman conducted dynamic centrifuge experiments on
concrete (1982) and aluminum (1985) cantilever retaining walls with dry dense sand
backfill and measured dynamic forces in agreement with the values predicted by the
M-O method, but suggested that the point of application should be located at mid-
height of the wall. More recently, Stadler (1996) performed a series of centrifuge
model experiments on cantilever walls founded on a rigid foundation and observed
that the total lateral earth pressure distribution was approximately triangular, while
dynamic earth pressure distribution varied between triangular and rectangular.
However, the magnitude of the forces was less than would be predicted using the M-
O method and the experimentally measured moments were on the order of 20 to 75%
less than would be predicted using the M-O method. Since Stadler’s model walls
were mounted directly on the base of the container, as was the case with most of the
above mentioned centrifuge studies, they shared some of the limitations of placing the
model on “bedrock” inherent in the 1-g shaking table experiments, albeit with much
more rigorous model scaling.
A cantilever wall with dry medium-dense sand backfill founded on soil was
examined in a series of dynamic centrifuge experiments by Ortiz et al. (1983). They
also observed a broad agreement between the maximum measured forces and the M-
O predictions; however, they found that the maximum dynamic force acted at about
one third the height of the wall above its base. Similar conclusion regarding the point
of application of the maximum dynamic force was reached by Al-Atik and Sitar
(2010) who also modeled retaining structures on soil foundation, although they found
that, in general, the loads predicted with the M-O method exceeded those measured in
the centrifuge experiments.
Gravity walls with dry medium-dense sand backfill on soil foundation were
studied by Nakamura (2006). He concluded that contrary to the M-O rigid wedge
assumption, the part of the backfill that follows the displacement of the retaining wall
deforms plastically while sliding down. Also, while the M-O theory assumes that no
phase difference occurs between the motion of the retaining wall and backfill,
Nakamura (2006) observed that the acceleration is transmitted instantaneously
through the retaining wall and then transmitted into the backfill and, therefore, the
dynamic earth pressures and inertia forces are not in phase. Most, importantly, while
the wall as a whole tended to translate laterally, the dynamic increment in load did not
significantly exceed the static load and the earth pressure remained roughly triangular
with depth.

ANALYSIS AND DESIGN

Mononobe – Okabe (M-O) Method and Its Derivatives for Yielding Walls. The
M-O method is based on the work of Okabe (1926) and Mononobe and Matsuo
(1929) following the great Kanto Earthquake of 1923 in Japan. It was originally
developed for gravity walls retaining cohesionless backfill materials and it is today

7
the most common approach to determine seismically induced lateral earth pressures
on a variety of structures.
This method uses a pseudo-static analysis based on the Coulomb wedge
theory for active and passive earth pressure and includes additional vertical and
horizontal seismic forces, as shown in Figure 6. The inherent assumptions in the
method are as follows:
 The wall yields sufficiently to produce minimum active pressures
 When the minimum active pressure is attained, a soil wedge behind the wall is
at the point of incipient failure and the maximum shear strength is mobilized
along the potential sliding surface
 The soil behind the wall behaves as a rigid body

i
kvW
khW
 
W


H/3 PA
AE

Figure 6 Forces Considered in the Mononobe-Okabe Analysis.

Using force equilibrium the total active thrust PAE per unit length of wall is
determined by:

PA E  0.5H 2 (1  kv )K A E (1)

Where,

cos2 (     )
K AE  2
(2)
 sin(   )  sin(     i ) 
cos   cos2   cos(     )  1  
 cos(     )  cos(i  ) 

and γ = unit weight of the soil, H = height of the wall,  = angle of internal friction of
the soil, δ = angle of wall friction, β = slope of the wall relative to the vertical, θ =
tan-1(kh/(1-kv)), kh = horizontal acceleration (in g), and kv = vertical acceleration (in
g).

8
Equation (1) represents the total active thrust on the wall during seismic
loading and the point of application of the resulting force is at 1/3H. As already
mentioned, a serious limitation of equation (2) is that it increases exponentially and
does not converge if θ <  − β (e.g. Kramer, 1996), which for typical values of angle
of internal friction means accelerations in excess of 0.7 g. The approach adopted in
such cases tends to be to use the simplified equation introduced by Seed and Whitman
(1970) that separated the total force on the wall into a static and dynamic component.
They then proposed a simplified expression for the dynamic increment of the active
thrust as:
PA E  1 2 H 2 K A E and K A E ~ 3 4 kh (3)
where kh is the horizontal ground acceleration as a fraction of the acceleration of
gravity. This approximation is asymptotically tangent to the M-O solution at
accelerations below about 0.4 g and it remains linear throughout. In addition, Seed
and Whitman (1970) recommend that the resultant of the dynamic force increment be
applied at 0.6H, hence introducing the concept of the “inverted triangle” to the
distribution of the dynamic force increment.
0 5 10 15 20 25
0
Pressure distribution
during shaking with
10 max. accn. =0.3g
Depth - cm

20

30
Initial static
pressures
40
Earth Pressure - g/cm2

Figure 7. Total (dynamic plus static) vs. static earth pressure (Seed and
Whitman (1970) after Matuso, 1941)

The impetus for the recommendation by Seed and Whitman (1970) to place
the resultant of the dynamic force increment at 0.6H is their interpretation of
Matsuo’s (1941) experimental results as reproduced in Figure 7. As stated earlier,
Mononobe and Okabe considered that the total pressure computed would act at a
height of H/3 above the base wall. However, a consensus has not been reached
regarding the appropriate point of application of the dynamic force increment as can
be seen from Table 1, which lists the recommendations of various researchers whose
earth pressures agree more or less well with the M-O earth pressures.

9
Table 1: Recommended Point of Application of The Seismic Force Increment

Author Point of Application


Mononobe-Okabe (1926-1929) 0.33H
Seed and Whitman (1970) 0.6H
Nandakumaran and Joshi (1973) <0.65H
Krishna et al. (1974) ~0.5H
Sherif et al. (1982) ~0.42H
Prakash and Brasavanna (1969) varies with acceleration
Ichihara and Matsuzawa (1973) varies with acceleration
Ortiz et al. (1983) varies, but higher than H/3
Woodward and Griffiths (1992) varies with acceleration
Steedman and Zeng (1990) varies, but higher than H/3
Mylonakis et al. (2007) 0.33H

Whereas the position of the point of application of the resultant has been the
subject of a continuing discussion, many researchers (performing both analytical and
experimental studies) have agreed that the earth pressures determined by the M-O
method for “yielding walls” give adequate results (e.g. Matsuo (1941), Ishii et al.
(1960), Prakash and Basavanna (1969), Seed and Whitman (1970), Ichihara and
Matsuzawa (1973), Clough and Fragaszy (1977), Bolton and Steedman (1982), Sherif
et al. (1982), Ortiz et al. (1983), Musante and Ortigosa (1984), Ishibashi and Fang
(1987), Steedman and Zeng (1990), Woodward and Griffiths (1992), Anvar and
Ghahramani (1995), Richards et al (1999)). Most recently, Mylonakis et al. (2007)
proposed a modification of the M-O method which results in a single equation for
both, active and passive condition. Their active solution is more conservative and
gives a total force somewhat higher than M-O, whereas their solution for the passive
case is less conservative than the M-O solution. They recommend 1/3H as the point
of application of the total force resultant.
Given the above, it is important to reiterate that Seed and Whitman (1970)
suggested that a “yielding” wall designed adequately for static forces will probably
perform well under seismic loading. This suggestion has been supported by Clough
and Fragaszy (1977) who proposed that a cantilever retaining structure with an
adequate factor of safety for static forces will have enough reserve to resist dynamic
loads up to an acceleration of at least 0.4g. Similarly, Whitman (1991) states that:
“Structures away from waterfronts have generally fared well during
earthquakes. Examples of stability-type failures are rare…”
The same sentiment is echoed by Huang (2000) who noted that failures of
gravity walls are generally due to bearing capacity failures and not due to substantial
increase in earth pressures. Finally, the commentary to ATC 32 states that most free-
standing retaining walls not associated with other structures have performed well
during past earthquakes even though no particular seismic design was implemented,
which is consistent with the observations noted above and with the authors’ own
experience.
Not everyone agrees with this point of view. Richards and Elms (1979) argued
that the M-O approach would lead to an unconservative estimate of the dynamic

10
thrust because the inertia of the wall is not considered. They suggested a new design
approach based on allowable displacements. This method based on Newmark’s
(1965) approach for determining seismically induced displacements in dams and
embankments was later reviewed and recommended by Whitman and Liao (1984).
More recently, Steedman and Zeng (1996) and Zeng and Steedman (2000) present a
displacement based method for the analysis of gravity walls using the same
principles. Nevertheless, currently, the M-O method remains the most widely used
method for “yielding” retaining walls.

Dynamic Earth Pressures on Non-Yielding/Rigid Walls. The currently accepted


tenet is that seismically induced earth pressures on rigid walls, such as basement
walls, are higher than those on yielding walls. This is of course true for the static case
where earth pressures for yielding walls are performed with the active earth pressure
coefficient KA, whereas the at-rest earth pressures K0 are used for unyielding walls.
Theoretical results obtained by Wood (1973) are still considered as a standard for this
case and the total dynamic thrust is approximated as ΔPAE= γH2A and the resultant
acting at 0.6H above the base (FEMA 750). This results in a total thrust and
corresponding moment approximately 2-3 times higher than that predicted by Seed
and Whitman (1970) for yielding walls. The solution assumes a rigid wall being acted
upon by elastic soil connected to a rigid foundation (Figure 8).

Homogenous elastic soil


y,v y = 0 (Plain strain) u=0
xy = 0 xy = 0

u=0
xy = 0 Rigid
 Uniform body H wall
force
u=0
v=0

Rigid boundary x,u


L

Figure 8. Geometry and boundary conditions assumed by Wood (1973).

FEMA 750 interpretation of the above condition is as follows:


“Most nonyielding walls will be located on rock or very stiff soil. Even in this
condition, wall flexibility can be sufficient to develop active seismic earth
pressures significantly reducing the loading on basement walls. Where a
basement wall is located on rock or very stiff soil and where structural
analyses determine that the wall flexibility is such that deformations will not

11
develop seismic active earth pressures (i.e., deformations < 0.002H where H
is the wall height), the wall should be designed as a nonyielding wall.”
Note that the assumed geometry is more akin to a rigid box constraining
elastic infill, as opposed to soil acting as an external load on a structure. In spite of
these unique boundary conditions assumed by Wood (1973), this approach has been
adopted by other researchers and similarly high pressures for unyielding walls were
obtained by Matsuo and Ohara (1960), Scott (1973), Prakash (1981), Sherif et al.
(1982), Veletsos and Younan (1997), Zhang et al. (1998), Ostadan and White (1998)
and Ostadan (2005).
Ostadan (2005) (also Ostadan and White, 1998) proposed a simplified method
which incorporated the main parameters affecting the seismic loading on basement
walls and Ostadan’s work is currently recommended by NEHRP (FEMA 750)
provisions for seismic regulations of structures. His analyses show that the
application of this method yields results that correspond in magnitude to the M-O
method as a lower bound and the Wood solution as the upper bound which is as much
as 2 to 2.5 times greater than the M-O solution.
However, if loads of this magnitude were to have been experienced by
basement walls we would expect to have seen evidence of damage to deeply
embedded structures/basements in recent earthquakes with very strong ground
motions in heavily developed areas which has not been the case (Lew et al. 2010a,
2010b). Whitman (1991) addressed this apparent inconsistency by suggesting that
because basements move relative to the foundation soil owing to soil structure
interaction, dynamic earth pressures equal to those obtained by M-O are adequate
except for a special condition of structures founded at a sharp interface between soil
and rock. Similarly, Veletsos and Younan (1997) and Younan and Veletsos (2000)
suggest that a substantial decrease from the typical “rigid” solutions is usually
obtained for walls of realistic flexibilities. Most recently, Ahmadnia et al. (2011)
performed a series of numerical analyses assuming typical basement wall
configurations including cross struts and found that the M-O method, as simplified by
Seed and Whitman (1970), gave adequately conservative results, with the point of
application of the dynamic force at about 0.5 H.

Elastic Wave Propagation Methods. One alternative to the limit equilibrium


method of analysis used in the M-O analysis is to consider elastic wave propagation.
Based on Zeng’s (1990) dynamic centrifuge experiments, Steedman and Zeng (1990)
suggested that the dynamic amplification or attenuation of input motion through the
soil and phase shift are important factors in the determination of the magnitude and
the distribution of dynamic earth pressures. In their derivation they assumed the same
Coulomb wedge as in the M-O method and they did not consider the reflected wave
at the ground surface, as illustrated in Figure 9. The consequence of this assumption
is that the computed earth pressure response is not amplified near resonance as shown
in Figure 10. Their solution gives results which are equal or lower than the M-O
method and suffers from the same limitations as the M-O method at high
accelerations. It is important to note that the results in Figure 10 are for specific
material properties in order to obtain the soil period, in this case φ = 33°, δ = 16°, Vs
= 109 m/s, Gmax= 20 GPa, and H=10m. In addition, the computed earth pressure

12
distribution at or below resonance i.e. the soil period matching the input motion
period, is triangular, increasing with depth, and identical to that assumed in the M-O
method. The resulting dynamic force increment is applied in the range 0.33 to 0.4H
from the base of the wall for most typical conditions.

Steedman & Zeng


Full wave equation Free Field

Qh
W
H

R
 P
AE
 Rock

utt (H) = k h gsin(t)

Figure 9. Geometry of the problem as defined by Steedman and Zeng (1990).


1
Steedman & Zeng
Wave Equation, =10%
0.8

0.6
k h=0.3g
Kae

k h=0.2g
0.4 k =0.1g
h
k h=0.05g
0.2

0
0 0.5 1 1.5 2 2.5 3 3.5 4
4H/Vs / T

Figure 10. Dynamic earth pressure increment as a function of the ratio of the
natural period of the soil deposit and the input motion period.

Effect of Cohesion. All of the above mentioned analytical solutions assume ideal
cohesionless backfill. Such conditions occur rarely and in such cases high water table
and liquefaction susceptibility often drive the problem. In most typical cases, the
natural deposits exhibit some degree of cohesion (see Figure 1) and, therefore, are
unlikely to mobilize the kind of forces predicted assuming ideal cohesionless backfill.
The contribution of cohesion to the reduction of seismic earth pressure on retaining
structures is explored by Anderson et al. (2008) who present a series of example
charts to illustrate the reduction in computed seismic coefficient as a function of
cohesion. They conclude that the “reduction for typical design situations could be on
the order of about 50 percent to 75 percent” and ascribe the good observed seismic
performance of retaining structures in part to the presence of cohesion in typical

13
backfill. Similar conclusions are reached by Lew et al. (2010a,b) based on
observations of seismic performance of different types of retaining structures in
recent earthquakes.

DATA FROM RECENT EXPERIMENTS

The preceding discussion exposed some of the dilemmas facing researchers and
practitioners in trying to determine the best approach to adopt in dealing with
seismically induced earthpressures based on the results of previous experimental
studies and field observations. To address some of these issues the authors are in the
midst of an experimental program to evaluate seismically induced earth pressures on
different types of retaining structures and to-date 3 dynamic centrifuge experiments
with 3 different configurations of retaining structures and backfill were performed, as
follows: 1) two identical U-shaped structures with cross bracing and medium dense
sand backfill; 2) a cantilever U-shaped structure and a free standing cantilever wall
with medium dense sand backfill ; and, 3) a cross-braced U-shaped structure and a
free standing cantilever with compacted low plasticity silty clay (Yolo Loam)
backfill.

Centrifuge Model Configurations. The centrifuge experiments were performed at


the Center for Geotechnical Modeling at the University of California, Davis, using the
flexible shear beam container. The centrifugal acceleration used in the experiments
to-date has been 36g and all test results are presented in terms of prototype units
unless otherwise stated. The experiments with the 6.5 m high, U-shaped cantilever
walls with sand backfill followed the same model construction procedures and used
the same model structures as used by Al-Atik and Sitar (2010) in order to allow a
direct comparison of the results. The same aluminum model structures were used and
lead was added to the structures to match the mass of the prototype reinforced
concrete structures. The structures were fully embedded in dry sand backfill with
relative density of 72% and were underlain by approximately 12.5m of sand or clayey
silt in prototype scale.
The basement type (rigid structures) were modeled by modifying the U-
shaped models by adding two levels of struts. The struts were instrumented with load
cells in order to obtain direct measurements of the dynamic loads on the walls as
shown in Figure 11. The free standing cantilever model was based on scaling of a 6.5
m prototype of California Department of Transportation design cantilever wall
(Caltrans, 2010). This design incorporates a shear key in the footing in order to limit
potential sliding of the footing during seismic loading. The silty clay (Yolo Loam)
has PI = 11% and LL = 30%. It was hand compacted at a water content 2% above
optimum to a relative compaction of 90% of Standard Proctor. Figure 12 is a diagram
showing the configuration of the experiment with compacted clay silt which included
one free standing cantilever wall and a cross-braced U-shaped structure. Scaled
earthquake records developed by Al Atik and Sitar (2010) were used in order to
maintain consistency for the purposes of comparison of the results. Multiple shaking
events covering a wide range of predominant periods and peak ground accelerations
were applied to each model in flight.

14
Figure 11. Photograph of the cross-braced, U-shaped structure model with sand
backfill. Note the load cell on each strut.

Figure 12. Schematic of the layout of the 3rd experiment showing the positions of
different transducers and the geometry of the models. All dimensions are in mm.

The models were instrumented to measure accelerations, displacements,


bending moments and earth pressures. Soil settlement and the deformation and
settlement of the structures were measured at different locations using a combination
of spring loaded LVDTs and linear potentiometers. The lateral earth pressures were
measured directly using flexible Tactilus™ pressure sensors. Lateral earth pressures
on the cantiler structures were also calculated by double differentiating bending
moments measured by the strain gages mounted on the model walls.

Dynamic Earth Pressures. As already discussed, the issue of the distribution and
magnitude of the seismically induced earth pressures is very important since it bears

15
directly on the computed forces and, most importantly, from a structural design
perspective, it directly affects the magnitude of the computed shear and moment
demand on the structure. Figures 13 and 14 show the measured distribution and
magnitude of the dynamic earth pressure increment for structures retaining
cohesionless and cohesive backfill, respectively, subjected to filtered and scaled
Kobe-TAK 090 input motion (Al Atik and Sitar, 2009). The corresponding dynamic
earth pressure increment obtained from the M-O method and from the Seed and
Whitman (1970) solution are plotted for comparison. All plots correspond to the time
of maximum moment on the respective structures.

Kobe TAK090
1 1 1

0.8 0.8 0.8

0.6 0.6 0.6


z/H

0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
qAE / H qAE / H qAE / H

M-O Load cells Pressure Cells


Seed & Whitman (1979) Strain gages
Figure 13. Dynamic earth pressure increment from experiments with medium
dense sand backfill in response to scaled Kobe-TAK090 input motion.

Kobe TAK090
1 1

0.8 0.8

0.6 0.6
z/H

0.4 0.4

0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
qAE / H qAE / H
M-O Load cells
Seed & Whitman (1979) Strain gages

Figure 14. Dynamic earth pressure increment from experiments with compacted
silty clay backfill in response to scaled Kobe-TAK090 input motion.

16
Three different means of obtaining earth pressure data were used. Tactile
sensors, denoted by solid dots, provided a direct measurement of the earth pressure.
However, their dynamic response is too slow to fully capture the peak response and,
therefore, they underestimate the magnitude of the pressure increment. Thus, they are
suitable only for defining the overall trend of earth pressure distribution. On the other
hand, the strain gauges and the load cell have excellent dynamic range, however, the
actual stress distribution is based on fitting a cubic polynomial to the deflected shape
of the structures as obtained by the strain gauges. The resulting earth pressure
distributions, denoted by the open dots, are linear.
Overall, the data show that the seismic earth pressure increments increase with
depth consistent with static earth pressure distribution and consistent with the M-O
solution as the upper bound for the experimental results. The “inverted triangle”
dynamic earth pressure increment, as computed using the Seed and Whitman (1970)
solution, while producing a reasonable magnitude of the maximum dynamic earth
pressure increment, does not reflect the observed distribution of the dynamic earth
pressure increment. These results are quite consistent with the results previously
obtained by Ortiz et al. (1983), Stadler (1996) and Al-Atik and Sitar (2010).
Figures 15, 16 and 17 are plots of the experimental results in terms of the
dynamic earth pressure coefficient, Kae, together with the curves obtained using the
most common analytical solutions against free field PGA at the point of maximum
measured moment on the respective structures. The plots show that the experimental
data exhibit considerable scatter with increasing PGA and duration of the input
ground motion. The scatter is particularly large for the models with the U-shaped
cantilever structures and is most likely related to a combination of factors, including
slight variations in relative density of the models and possible boundary effects that
become more pronouced at higher accelerations.
1.2
M-O
Seed & Whitman , 1970
1 Mylonakis et al., 2007
Al Atik & Sitar, 2010
Current data (sand)
0.8
Kae

0.6

0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Free Field PGA

Figure 15. Seismic earth pressure coefficient as a function of PGA for U-shaped
cantilever walls with medium dense sand backfill.

17
1.2
M-O
Seed & Whitman , 1970
1 Wood, 1973
Mylonakis et al., 2007
Current data (sand)
0.8 Current data (cohesive soil)
Kae
0.6

0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Free Field PGA

Figure 16. Seismic earth pressure coefficient as a function of PGA for cross-
braced walls.

1.2
M-O
Seed & Whitman , 1970
1 Mylonakis et al., 2007
NCHRP c/H = 0.15
Current data (sand)
0.8 Current data (cohesive soil)
Kae

0.6

0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Free Field PGA

Figure 17. Seismic earth pressure coefficient as a function of PGA for free
standing cantilever walls.
Nevertheless, the overall trends in the data show that the Seed and Whitman
(1970) approximation represents a reasonable upper bound for the value of the
seismic earth pressure increment for both fixed base cantilever structures (U-shaped
walls, Figure 15) and cross-braced, basement type, walls (Figure 16). In comparison,
the M-O solution and the Mylonakis et al. (2007) solution are considerably higher
than measured values at accelerations above about 0.4 g. The equivalent Wood (1973)
seismic coefficient, computed using the prototype structure dimensions, clearly
exceeds all other results by a considerable margin as would be expected based on the
assumptions used in deriving this solution, as discussed earlier.

18
The most significant difference between the analytically predicted seismic
earth pressure increment and the observed data is for the free standing cantilever
walls. The fact that small amount of rotation and translation can significantly
decrease the forces acting in a retaining structure have been well recognized (e.g.
Anderson et al., 2008) and the data presented in Figure 17 clearly shows this to be the
case. In order to arrive at more moderate seismic earth pressures it is a common
practice to include a small amount of cohesion (Anderson et al. 2008). The curve for
cohesive backfill plotted in Figure 17 corresponds to the material properties of the
compacted silty clay used in the experiment. As the plot show this is still a very
conservative result. Overall, this data is consistent with the position presented by
Seed and Whitman (1970) who suggested that well designed retaining structures
should be well capable of withstanding ground motions with PGA on the order of
0.3g without the need for specific seismic design.

Dynamic Moments. Dynamic moments are ultimately the quantities that dictate the
structural design of the retaining structures. Clearly, the magnitude of the seismic
earth pressure increment is very important in this regard. However, even more
important is the point of application of the resultant, since the decision whether to
apply the resultant at 0.6H versus 1/3H immediately changes the computed moment
by a factor of about 2. The significance of this effect is illustrated in Figure 18
showing the dynamic moment increment plotted against PGA for the case of a free
standing cantilever retaining wall. These results show that the M-O method gives
amply conservative results over the full range of accelerations and that applying the
seismic earth pressure increment at 0.6H, as recommended by Seed and Whitman
(1970) and many others, leads to a significant, if not unnecessary, overdesign.

0.2
M-O
Seed & Whitman , 1970
Mylonakis et al., 2007
0.16
NCHRP c/H = 0.15
Current data (sand)
Current data (cohesive soil)
0.12
3
Mae / H

0.08

0.04

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Free Field PGA

Figure 18. Maximum dynamic moment increment as a function of PGA for free
standing cantilever walls.

19
DISCUSSION AND CONCLUSIONS

The review of the previously developed analysis and design procedures and of the
corresponding experimental data shows a significant difference in the observed and
perceived distribution of the seismic earth pressure increment. The most likely
difference between the different experimental results appears to be in the execution of
the experiments. Specifically, 1-g shaking table experiments suffer from significant
scaling and boundary condition limitations that cannot be easily overcome. The most
significant limitation is the rigid base of the models built directly on the shaking
table. In this respect, some of the past centrifuge experiments share the same
limitation and it is the data from these experiments, starting with the work of
Mononobe and Matsuo (1929) that suggest that seismic earth pressures increase
toward the ground surface. In contrast, the seismic centrifuge experiments in which
the structures are based on a soil foundation, show the opposite trend, with the
seismic earth pressure increment increasing downward. These results point to the
need to carefully evaluate the suitability of experimental facilities, especially 1-g
shaking tables, for modeling of structures embedded in soil or based on soil
foundation.
The most direct impact of the recognition that the point of application of the
seismic earth pressure increment can be reasonably placed at 1/3H is the reduction in
the computed design moments for the structure. Another important aspect of the
results presented herein is the observation that stiff embedded structures do not seem
to experience substantial increase in seismic earth pressure over that experienced by
cantilever structures with fixed base. In this regard the Wood (1973) solution is not
representative of conditions commonly encountered in practice and its continued use
is not recommended.
The new experimental data suggest that the simplified M-O method proposed
by Seed and Whitman (1970) provides an ample and reasonable upper bound for the
expected magnitude of seismic earth pressure increment for moderate height retaining
structures, 6-7 m high, in level ground, recognizing that the resultant force should be
applied at 1/3H. The same applies to basement walls or cross-braced excavations.
However, caution should be used in extrapolating these results to deeper embedded
structures. There is no theoretical basis that would lead one to expect that seismic
earth pressures continue to increase monotonically with depth. At some point the
structures will begin to move with the soil and their response will become more
consistent with that of tunnels which are well known to perform very well under
seismic loading. An additional limitation of the experimental results presented herein
is that they apply only to level ground. However, retaining structures are frequently
placed on slopes with sloping backfill and sloping ground below. This type of setting
requires a different approach, as slope stability, rather than earth pressure may be the
governing mechanism of failure. At present this distinction frequently is not made.
All these issues deserve further careful evaluation, since the costs of an over-
conservative design can be just as much of a problem as the cost of a future failure. In
this respect, while there is a need for further experimental work, there is a much
greater need for the development of a database of field observations from
instrumented sites and structures. Only then we will be able to evaluate fully the

20
actual performance under a variety of conditions. The development of such data is
essential if we are to advance the state of the art and take the full advantage of
advanced analytical tools that go hand in hand with modern performance based
design.

ACKNOWLEGMENTS

The experimental program carried out in this research could not have been executed
without the able assistance of Nathaniel Wagner and Jeff Zayas of the University of
California at Berkeley. Dr. Dan Wilson and the staff of the Center for Geotechnical
Modeling at UC Davis have been most accommodating and provided outstanding
environment for a truly collaborative effort. The research funding was provide in part
by a grant from the California Geotechnical Engineering Association (CalGeo), the
State of California Department of Transportation (Caltrans) Contract No. 65N2170
and NSF-NEES-CR Grant No. CMMI-0936376: Seismic Earth Pressures on
Retaining Structures. G. Candia was funded in part by a fellowship from the Chilean
Commission for Scientific Research and Technology (CONICYT).

REFERENCES

Ahmadnia A., Taiebat M., Finn W.D.L., Ventura, C.E. and Devall, R.H. (2011).
“Seismic Assessment of Basement Walls for Different Design Criteria,”
Proceedings of 2011 Pan-Am CGS Geotechnical Conference, Toronto, Ont.
Anderson, D.G., Martin, G.R., Lam, I.P. and Wang, J.N. (2008). “Seismic Design and
Analysis of Retaining Walls, Buried Structures, Slopes and Embankments”,
NCHRP Report 611. Transportation Research Board, National Cooperative
Highway Research Program, Washington, D.C.
Anvar, S.A. and Ghahramani, A. (1995). “Dynamic Active Earth Pressure by Zero
Extension Line”, Proc. Third Int. Conf. on Recent Advances in Geotechnical
Earthquake Engineering and Soil Dynamics, St. Louis, Missouri, USA.
Aitken, G.H. (1982). Seismic Response of Retaining Walls, MS Thesis, University of
Canterbury, Christchurch, New Zealand.
Al Atik, L. and Sitar, N. (2009). Experimental and Analytical Study of the Seismic
Performance of Retaining Structures. Pacific Earthquake Engineering Research
Center, PEER 2008/104, Berkeley, CA.
Al Atik, L. and Sitar, N. (2010), "Seismic Earth Pressures on Cantilever Retaining
Structures," Journal of Geotechnical and Geoenvironmental Engineering,
October, (136) 10, pp. 1324-1333.
Bolton M.D. and Steedman, R.S. (1982). “Centrifugal Testing of Micro-Concrete
Retaining Walls Subject to Base Shaking,” Proceedings of Conference on Soil
dynamics and Earthquake Engineering, Southampton, 311-329, Balkema.
Bolton M.D. and Steedman, R.S. (1985). “The Behavior of Fixed Cantilever Walls
Subject to Lateral Loading,” Application of Centrifuge Modeling to Geotechnical
Design, Craig (ed.), Balkema, Rotterdam.

21
Building Seismic Safety Council. (2004). NEHRP Recommended Provisions for
Seismic Regulations for New Buildings and Other Structures (FEMA 450), 2003
Edition, Part 1 – Provisions. BSSC, Washington, DC.
Building Seismic Safety Council. (2010). NEHRP Recommended Provisions for
Seismic Regulations for New Buildings and Other Structures (FEMA 750), 2009
Edition, Part 1 – Provisions. BSSC, Washington, DC.
California Department of Transportation (Caltrans). (2010). “Standard Plans:
Retaining Wall Type 1 - H = 4' through 30', Plan No. B3-1”, Standard Plans- U.
S. Customary Units, Department of Transportation State of California.
Clough, G.W. and Fragaszy, R.F. (1977). “A Study of Earth Loadings on Floodway
Retaining Structures in the 1971 San Fernando Valley Earthquake,” Proceedings
of the Sixth World Conference on Earthquake Engineering, Vol. 3.
Dewoolkar, M.M., Ko, H. and Pak R.Y.S. (2001). “Seismic Behavior of Cantilever
Retaining Walls with Liquefiable Backfills,” Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, (127) 5, 424-435.
Gazetas, G, Psarropoulos, PN, Anastasopoulos, I. and Gerolymos, N. (2004).
“Seismic Behaviour of Flexible Retaining Systems Subjected to Short-Duration
Moderately Strong Excitation,” Soil Dynamics and Earthquake Engineering, (24),
537-550.
Huang, C. (2000). “Investigations of Soil Retaining Structures Damaged During the
Chi-Chi (Taiwan) Earthquake.” Journal of the Chinese Institute of Engineers, (23)
4, 417-428.
Ichihara, M. and Matsuzawa, H. (1973). “Earth Pressure During Earthquake”, Soils
and Foundations, (13) 4, pp. 75-86.
Ishibashi, I., and Fang, Y.S. (1987). “Dynamic earth pressures with different wall
movement modes,” Soils and Foundations, (27) 4, 11-22.
Ishii, Y. Arai, H. and Tsuchida, H. (1960). “Lateral earth pressure in an earthquake”,
In: Proceedings of the 2nd World Conference on Earthquake Engineering, Tokyo,
Vol. 1, pp. 211-230.
Kramer, S.L., (1996). “Geotechnical Earthquake Engineering”, New Jersey, Prentice
Hall.
Krishna, J., Prakash, S. and Nandkumran, P. (1974). "Dynamic Earth Pressure
Distribution Behind Flexible Retaining Walls," Journal, Indian Geotechnical
Society, (4) 3, pp. 207-224.
Lew, M., Sitar, N., and Al Atik, L. (2010a). “Seismic Earth Pressures: Fact or
Fiction.” Invited Keynote Paper, Earth Retention Conference, ER 2010, ASCE,
Seattle.
Lew, M., Sitar, N., Al Atik, L., Pourzanjani, M. and Hudson, M.B. (2010b). “Seismic
Earth Pressures on Deep Building Basements.” Structural Engineers Association
of California, Proceedings of the Annual Convention, 2010.
Matsuo, H. (1941). “Experimental Study on the Distribution of Earth Pressures
Acting on a Vertical Wall during Earthquakes,” Journal of the Japanese Society
of Civil Engineers, (27) 2.
Matsuo, H. and Ohara, S. (1960). “Lateral Earth Pressure and Stability of Quay Walls
During Earthquakes,” Proceedings, Second World Conference on Earthquake
Engineering, Vol. 1, Tokyo, Japan.

22
Mononobe, N. and Matsuo M. (1929). “On the Determination of Earth Pressures
during Earthquakes,” Proceedings, World Engineering Congress, Vol. 9, 179-
187.
Mylonakis, G., Kloukinas, P. and Papatonopoulos, C. (2007). “An Alternative to the
Mononobe-Okabe Equation for Seismic Earth Pressures”, Soil Dyn. and
Earthquake Eng., (27) 10, 957-969.
Musante, H. and Ortigosa P. (1984). “Seismic Analysis of Gravity Retaining Walls”,
Proceedings, Eighth World Conference on Earthquake Engineering, San
Francisco, USA.
Nakamura, S. (2006). “Reexamination of Mononobe-Okabe Theory of Gravity
Retaining Walls Using Centrifuge Model Tests,” Soils and Foundations, (46) 2,
135-146.
Nandakumaran, P. and Joshi, V.H. (1973). “Static and Dynamic Active Earth
Pressure behind Retaining Walls”, Bulletin of the Indian Society of Earthquake
Technology, (10) 3.
Nazarian, H.N. and Hadjian, A.H. (1979). “Earthquake Induced Lateral Soil Pressures
on Structures,” Journal of Geotechnical Engineering Division, ASCE, (105) GT9:
1049-1066.
Newmark, N.M. (1965). “Effects of Earthquakes on Dams and Embankments,” Fifth
Rankine Lecture, Geotechnique, (15) 2, 139-160.
Okabe S. (1926). “General Theory of Earth Pressure,” Journal of the Japanese
Society of Civil Engineers, Tokyo, Japan, (12) 1.
Ortiz, L.A., Scott, R.F., and Lee, J. (1983). “Dynamic Centrifuge Testing of a
Cantilever Retaining Wall,” Earthquake Engineering and Structural Dynamics,
(11): 251–268.
Ostadan, F. (2005). “Seismic Soil Pressure for Building Walls – An Updated
Approach,” Journal of Soil Dynamics and Earthquake Engineering, (25): 785-
793.
Ostadan, F. and White, W.H. (1998). “Lateral Seismic Soil Pressure-An Updated
Approach”. In Preproceedings of UJNR Workshop on Soil-Structures Interaction,
U.S. Geological Survey, Menlo Park, California.
Prakash, S. and Basavanna, B.M. (1969). “Earth Pressure Distribution behind
Retaining Wall during Earthquakes,” Proceedings of the Fourth World
Conference on Earthquake Engineering, Santiago, Chile.
Prakash, S. (1981). "Dynamic Earth Pressures," State of the Art Report - International
Conference on Recent Advances on Geotechnical Earthquake Engineering and
Soil Dynamics, St. Louis, Missouri, Vol. III, 993-1020.
Richards, R, and Elms, D.G. (1979). “Seismic Behavior of Gravity Retaining Walls,”
Journal of the Geotechnical Engineering Division, ASCE, (105) GT4: 449–64.
Richards, R., Huang, C. and Fishman, K.L. (1999). “Seismic Earth Pressure on
Retaining Structures,” Journal of Geotechnical and Geoenvironmental
Engineering, ASCE, 125(9): 771–778.
Scott, R.F. (1973). “Earthquake-induced Earth Pressures on Retaining Walls,”
Proceedings, Fifth World Conference on Earthquake Engineering, Vol. 2, Rome,
Italy.

23
Seed, H.B. and Whitman, R.V. (1970). “Design of Earth Retaining Structures for
Dynamic Loads,” ASCE Specialty Conference, Lateral Stresses in the Ground and
Design of Earth Retaining Structures, Cornell Univ., Ithaca, New York, 103–147.
Sherif, M.A. and Fang, Y.S. (1984). “Dynamic Earth Pressures on Walls Rotating
about the Top,” Soils and Foundations, Vol. 24, No. 4, 109-117.
Sherif, M.A., Ishibashi, I., and Lee, C.D. (1982). “Earth Pressure against Stiff
Retaining Walls,” Journal of Geotechnical Engineering, ASCE, 108, 679-695.
Stadler, A.T. (1996). “Dynamic Centrifuge Testing of Cantilever Retaining Walls,”
PhD Thesis, University of Colorado at Boulder.
Steedman, R.S. and Zeng, X. (1990). “The Seismic Response of Waterfront Retaining
Walls,” Design and Performance of Earth Retaining Structures, Conference
Proceedings, Cornell University, Ithaca, New York, June 18-21, ASCE
Geotechnical Special Publication No. 25.
Steedman, R.S. and Zeng, X. (1991). “Centrifuge Modeling of the Effects of
Earthquakes on Free Cantilever Walls,” Centrifuge’91, Ko (ed.), Balkema,
Rotterdam.
Steedman, R. S. and X. Zeng. (1996). “Rotation of Large Gravity Retaining Walls on
Rigid Foundations Under Seismic Loading,” in Analysis and Design of Retaining
Walls Against Earthquakes, ASCE/SEI Geotechnical Special Publication 60,
edited by S. Prakash, pp. 38-56.
Veletsos, A.S. and Younan, A.H. (1997). “Dynamic Response of Cantilever
Retaining Walls,” Journal of Geotechnical and Geoenvironmental Engineering,
(123) 2: 161-172.
Whitman, R.V. (1991). “Seismic Design of Earth Retaining Structures,” Proceedings,
Second International Conference on Recent Advances in Geotechnical
Earthquake Engineering and Soil Dynamics, ASCE, St. Louis, MO., 1767-1778.
Whitman, R.V. and Liao, S. (1984). “Seismic Design of Gravity Retaining Walls”,
Proceeding, Eight World Conference on Earthquake Engineering, San Francisco,
USA.
Wood, J.H. (1973). “Earthquake Induced Soil Pressures on Structures,” PhD Thesis,
California Institute of Technology, Pasadena, CA.
Woodward, P.K. and Griffiths, D.V. (1992). “Dynamic Active Earth Pressure
analysis,” Proc. 4th Int. Symp. on Numerical Models in Geomechanics (NUMOG
IV), Swansea, (eds. G.N. Pande and S. Pietruszczak), Pub. Balkema, Vol.1,
pp.403-410.
Younan, A.H. and Veletsos, A.S. (2000). “Dynamic Response of Flexible Retaining
Walls,” Earth. Eng. Struct. Dyn., (29):1815-1844.
Zhang, J.M., Shamoto, Y. and Tokimatsu, K. (1998). “Evaluation of Earth Pressure
under Lateral Deformation,” Soils and Foundations, (38)1: 15-33.
Zeng, X. (1990). “Modelling the Behavior of Quay Walls in Earthquakes,” PhD.
Thesis, Cambridge University, Cambridge, England.
Zeng X. and Steedman, R.S. (2000). “Rotating block method for seismic
displacement of gravity walls,” Journal of Geotechnical and Geoenvironmental
Engineering, ASCE, (126) 8: 709–717.

24

View publication stats

You might also like