You are on page 1of 19

Finite-Element Study for Seismic Structural and Global

Stability of Cantilever-Type Retaining Walls


Junied Bakr1; Syed Mohd Ahmad2; and Domenico Lombardi3

Abstract: This paper investigates the structural and global stability of a cantilever-type retaining wall under seismic loading using numerical
modelling. A new and robust approach is proposed to compute the seismic earth pressure behind the stem and along a virtual plane passing the
heel of the wall. The results show that under different earthquake characteristics and wall geometries, the seismic earth pressure forces may be
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

out of phase, leading to different seismic responses of the wall. The critical scenario for the structural stability is observed when the maximum
acceleration is directed toward the backfill soil, and the earthquake frequency content is close to the natural frequency of the wall. In contrast,
the critical scenario for the global stability occurs when the maximum acceleration is directed with minimum frequency content. Further, the
natural frequency of the wall does not affect the global stability of the wall. However, the duration of the applied earthquake acceleration does
affect the global stability of the wall, whereas the structural stability remains unaffected by it. In contrast with the current understanding, the
possibility of failure of a cantilever-type retaining wall by horizontal sliding is remarkably increased with time of the applied earthquake accel-
eration. DOI: 10.1061/(ASCE)GM.1943-5622.0001505. This work is made available under the terms of the Creative Commons Attribution
4.0 International license, http://creativecommons.org/licenses/by/4.0/.
Author keywords: Seismic stability analyses; Relative displacement; Wall; Backfill and stem seismic inertia forces; Shear force and bend-
ing moment; Seismic earth pressure increment.

Introduction In the current design practice, the cantilever-type retaining walls


are designed by considering the concepts used for the design of rigid
A cantilever-type retaining wall, like the one presented in Fig. 1(a), retaining walls; this is despite the important fact that the cantilever-
is one of the most common types of retaining structures. For such type retaining walls behave as flexible structures, whereas the rigid
walls, the most important design load that needs to be considered for retaining walls do behave as rigid structures. For a seismic case, the
a safe design under seismic conditions comes from the seismic earth force-based Mononobe Okabe (M-O) method (Mononobe and
pressure. For design purposes, a cantilever-type retaining wall is Matsuo 1929)—a method primarily based on Coulomb’s earth pres-
considered as a flexible structure and a design must address the sure theory (Coulomb 1776)—is widely used to estimate seismic
structural and global stability arising because of the seismic earth earth pressure for the retaining walls. Although the M-O method is
pressure. As presented in Fig. 1(b) the seismic earth pressure will a very straightforward method to use, it has a limitation of not repli-
create a shear force, Nwall, and bending moment, Mwall, on the stem cating, and hence simulating, the real field behavior and, therefore,
of the wall and also tend to cause the base slab to slide relative to the at times overestimating the seismic earth pressure (e.g., Nakamura
foundation layer and rotate about the toe, thereby overturning the 2006; Al Atik and Sitar 2009; Jo et al. 2017; Bakr and Ahmad
wall [Fig. 1(c)]. For structural stability, the stem of the wall should 2018a). Extensive efforts have been made to study the seismic earth
be designed to resist the shear force, Nwall, and bending moment, pressure behind a rigid retaining wall (e.g, Choudhury and Katdare
Mw, whereas, for global stability, the wall should be designed such 2013; Tang et al. 2014; Dey et al. 2017; Zhou et al. 2018; Rajesh
that it provides adequate sliding and overturning resistance. It is cru- and Choudhury 2017), whereas little attention has been paid to
cial to highlight that for both of these stability analyses assessments, investigate the development of seismic earth pressure behind a
an accurate estimation of the seismic earth pressure is important. cantilever-type retaining wall. Further, with regards to finding a
critical case that needs to be considered for designing a cantilever-
type retaining wall under seismic conditions, there are several con-
1 tradictory evidences in the available literature. For example, the nu-
Formerly, Ph.D. Student, School of Mechanical, Aerospace and Civil
merical study presented by Green et al. (2008) and the shaking table
Engineering, Univ. of Manchester, Manchester M13 9PL UK. ORCID:
https://orcid.org/0000-0003-1441-0425. study presented by Kloukinas et al. (2015) show that a critical load-
2
Lecturer in Geotechnical Engineering, School of Mechanical, ing case for the seismic design of a cantilever-type retaining wall is
Aerospace and Civil Engineering, Univ. of Manchester, Manchester when the earthquake acceleration is applied away from the backfill
M13 9PL UK (corresponding author). Email: mohammed.ahmad.syed@ soil, thereby rendering the retaining wall–soil system in a passive
manchester.ac.uk earth pressure condition. This, however, is contradicted by the
3
Lecturer in Geotechnical Engineering, School of Mechanical, centrifuge-based study of Jo et al. (2017), who reported that a criti-
Aerospace and Civil Engineering, Univ. of Manchester, Manchester cal loading case for the retaining wall will be the one in which the
M13 9PL UK. ORCID: https://orcid.org/0000-0003-4116-8347.
earthquake acceleration is applied toward the backfill soil, thereby
Note. This manuscript was submitted on June 19, 2018; approved on
April 12, 2019; published online on August 16, 2019. Discussion period rendering the retaining wall–soil system in an active earth pressure
open until January 16, 2020; separate discussions must be submitted for condition. Cakir (2013) investigated the effect of frequency content
individual papers. This paper is part of the International Journal of of earthquake acceleration on the seismic response of a cantilever
Geomechanics, © ASCE, ISSN 1532-3641. retaining wall in a three-dimensional space by using the finite-

© ASCE 04019117-1 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


element method (FEM). This study found that the stresses behind pressure on the structural and global stability of a cantilever-type
the retaining wall and displacements at the top of the wall tend to retaining wall–soil system. The study also captures the deforma-
increase with the decrease of frequency content of earthquake accel- tion mechanism of a cantilever-type retaining wall–soil system
eration. Considering the importance of displacement-based meth- and investigates the effect of the natural frequency of a cantilever-
ods, researchers have presented various approaches for the estima- type retaining wall–soil system on its structural and global stability.
tion of displacement of rigid retaining walls (e.g., Nadim and Recommendations that would help in providing a safer and eco-
Whitman 1983; Madabhushi and Zeng 1998; Zeng 1998; Zeng and nomic design of a cantilever-type retaining wall for earthquake
Steedman 2000; Paruvakat et al. 2001; Huang 2005; Choudhury prone areas are also presented.
and Nimbalkar 2008; Basha and Babu 2010; Huang et al. 2009;
Ahmad and Choudhury 2010; Ni et al. 2017). Similarly, a few
researchers like Green et al. (2008); Kloukinas et al. (2015); Bakr Seismic Structural and Global Stability of a
and Ahmad (2018b) have investigated the deformation mechanism Cantilever-Type Retaining Wall–Soil System
of the cantilever-type retaining wall under the effect of seismic
loading. However, very little information is available that details a A cantilever-type retaining wall of height H and width of base slab
relationship and effect of seismic earth pressure on the displacement bslab, supporting a horizontal surface backfill to its full height, is pre-
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

of the cantilever-type retaining wall, thus leaving this important sented in Fig. 1(a). As mentioned above, the retaining wall–soil sys-
area of research and application poorly understood. tem is to be analyzed for the seismic structural and global stability.
This paper presents a FE-based numerical modeling approach
to study the seismic structural and global stability of a cantilever-
Seismic Structural Stability
type retaining wall–soil system, not only for a post-earthquake
scenario, but also during an earthquake event. The present study For the seismic structural stability of the wall, the wall is considered
identifies a critical scenario for the structural and global stability to be subject to the seismic earth pressure force coming from the
of a cantilever-type retaining wall–soil system by considering dif- backfill soil, which is assumed to act behind the stem, and denoted
ferent earthquake characteristics and the effects of seismic earth as Pstem, and to the wall seismic inertia force, Fwa and Fwp; where

(a) (b) (c)

(d) (e)

Fig. 1. Typical cantilever-type retaining wall–soil system: (a) schematic representation; (b) schematic of structural stability; (c) schematic of global
stability; (d) body force diagram for structural stability analysis; and (e) body force diagram for global stability analysis.

© ASCE 04019117-2 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


Fwa and Fwp are the wall seismic inertia force acting away and any boundary effects. For the purpose of analysis and for comparison
toward the backfill soil, respectively [Fig. 1(d)]. For Pstem, Fwa, and of the results of the proposed model with the centrifuge test results of
Fwp, the following needs to be noted: Jo et al. (2014), the stem of the retaining wall has been considered to
• At a given point in time, Fwa and Fwp will not be acting simul- be of a uniform stiffness. No embedment has been considered in front
taneously—it will either be Fwa or Fwp. of the cantilever-type retaining wall. This is a slight deviation from
• Depending upon the direction of the applied earthquake accel- actual practice, but this simplification facilitates comparison of the
eration—that is, whether it is acting toward the backfill soil or present model with a centrifuge-based study of Jo et al. (2014), who,
away from it, the wall seismic inertia force will also either be in their experimental model, did not consider any embedment, and
acting toward the backfill soil, denoted as Fwp [Fig. 1(d)] or simplifies numerical modeling to concentrate on the soil behavior.
away from the backfill soil, denoted as Fwa [Fig. 1(d)]. Further, because of impartial mobilization of passive resistance, a no-
• Because the amplitude of the applied earthquake acceleration will embedment case does not lead to unrealistic results. For the FEM, the
vary with time, Pstem, Fwa, and Fwp will also be time-varying. backfill and foundation soil is modeled by using six-noded triangular
• Pstem, Fwp, and Fwa will produce shear force, Nwall, and bend- elements with 2 degrees of freedom at each node and three Gauss
ing moment, Mwall, on the stem of the wall. integration points, whereas the wall is modeled by using the plate ele-
This paper presents a methodology to accurately predict Nwall ments of the PLAXIS2D 2016 library (Brinkgreve et al. 2016).
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

and Mwall; to estimate the relative contributions of Pstem, Fwa, and


Fwp on Nwall and Mwall; and to identify a critical case for the seismic Wall–Soil Interaction
structural stability of the wall that causes a maximum load case for
the stem of the wall. The interaction between the stem of the wall and backfill soil, and
between the base slab of the wall and foundation soil, has been mod-
Seismic Global Stability eled by using the six-noded interface elements of the PLAXIS2D
2016 library (Brinkgreve et al. 2016). For the chosen interface ele-
For the seismic global stability analysis of the wall, the wall is con- ment, the interface roughness was controlled by using an interface
sidered to be subject to [Fig. 1(e)]: strength-reduction-factor, Rinter, where Rinter = 0 for a perfectly
• The seismic earth pressure force, coming from the backfill smooth interface, Rinter = 1 for a perfectly rough interface, and
soil, which is assumed to act at a vertical plane passing through 0 < Rinter > 1 for a partially rough interface. To simulate the inter-
the heel of the wall, and denoted as Pvp. face properties similar to the Jo et al. (2014) centrifuge tests, for the
• The backfill seismic inertia force, Fsa and Fsp, where Fsa and present study, a partially rough interface was considered, such that
Fsp are the backfill seismic inertia force acting toward and Rinter = 0.334 between the stem of the wall and backfill soil and 0.5
away from the wall, respectively. between the base slab of the wall and foundation soil.
• The wall seismic inertia force, Fwa and Fwp.
Like Fwa and Fwp, depending upon the direction of the applied
Boundary Conditions
earthquake acceleration, the backfill seismic inertia force will also
either be acting away from the wall, denoted as Fsp [Fig. 1(e)] or to- For the FE model, the lateral boundaries were restrained against any
ward the wall, denoted as Fsa [Fig. 1(e)]. In addition, at any given horizontal movement, whereas at the base boundary, both the hori-
point in time, the wall–soil system will be subject to either Fsp or zontal and vertical movements were restricted [Fig. 2(a)]. Further,
Fsa; that is, they will not both be acting simultaneously. It is impor- absorbing lateral boundaries were considered so that the effects of
tant to highlight that the seismic earth pressure force, Pvp, is com- reflection of the seismic waves were negligible for the boundaries.
puted along the virtual plane because the global stability of the
cantilever-type retaining wall is maintained, in addition to the weight Constitutive Models for the Backfill and Foundation
of the wall, Ww, by the weight of the backfill soil above the base Soil, and the Material of the Wall
slab, Ws. This paper presents a methodology to accurately predict
the deformation mechanism of the wall to compute the relative dis- During a seismic analysis, a soil depicts a strong nonlinear stress–
placement between its base slab and foundation soil; to estimate the strain behavior, which is accompanied with a gradual reduction in
contribution of Pvp to the aforementioned relative displacement; and the shear modulus, and a simultaneous increase in hysteretic damp-
to identify a critical case for the global stability of the wall. ing with an increase in the shear strain amplitude (Brinkgreve et al.
For both the seismic structural and global stability analyses, the 2016). Thus, it is imperative that the chosen constitutive model for
effects of applied earthquake acceleration characteristics (i.e., its the present study should replicate the aforementioned soil behavior.
amplitude and frequency content) and the natural frequency of the In PLAXIS2D 2016 (Brinkgreve et al. 2016), the best model that
wall–soil system have been studied in detail. can replicate all of the above aspects of the constitutive behavior of
a soil is the hardening-soil-with-small-strain model (denoted as
HSsmall model from this point hereafter in the present paper, and
Finite-Element Model of the Cantilever-Type the same model was chosen for modeling the backfill and founda-
Retaining Wall–Soil System tion soils. However, because the HSsmall model is almost linear for
a very small strain with no hysteretic damping that would cause
An FE model of the cantilever-type retaining wall–soil system has unrealistic resonance during a numerical analysis, Rayleigh damp-
been developed by using the PLAXIS2D 2016 software (Brinkgreve ing was used in the present study by adopting the procedure outlined
et al. 2016). To validate the FE model and compare the results, the ge- by Rajasekaran (2009), which suggests
ometry of the model and the material properties were chosen to be
similar to the one used by Jo et al. (2014) for their centrifuge tests. ½C ¼ a½M þ b ½K (1)
The FE model is presented in Fig. 2(a), in which the wall has a height
of 5.4 m and sits on 9-m-thick foundation soil. The wall retains a hori- where [C], [M], and [K] = damping, mass and stiffness FE matrices
zontal surface backfill soil to its full height, and the confines of the of the wall–soil system; and a and b = Rayleigh parameters, com-
model in the horizontal direction are large enough so as to exclude puted as (Rajasekaran 2009)

© ASCE 04019117-3 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

(a)

(b) (c)

Fig. 2. (a) FEM of the cantilever-type retaining wall–soil system chosen for the present study; (b) 1989 Loma Prieta earthquake acceleration–time
history (data from PEER 2018); and (c) frequency content for (b).

( ) ( )
a 2 z s;wall v z1  v z2 vertical seismic inertia and, as a consequence, may slightly affect
¼ (2) the performance response of the cantilever-type retaining wall.
b v z1 þ v z2 1
However, arguments of several past studies (like Green et al. 2008;
Cakir 2013; Kloukinas et al. 2015; and Jo et al. 2017) as well as that
where z s,wall = damping ratio of the soil and wall, respectively; and of Bakr and Ahmad (2018a) suggest that it is primarily the horizon-
v z1 and v z2 = first two natural circular frequencies of the FE model. tal seismic acceleration that contributes to the seismic earth pressure
ABAQUS software (ABAQUS 2013) was used to compute the first force and affects the stability of a retaining wall. The same has been
two natural circular frequencies of the FE model, and for the model observed through the results of the present study as well as shown
presented in Fig. 2(a), they were found to be v z1 = 28.27 rad/s and later in the results section of this paper.
v z2 = 39.93 rad/s. The constitutive behavior of the material of the
wall was simulated by using a linear viscoelastic constitutive
model. Construction Sequence Simulation and Results from
the Static Analysis
Seismic Loading
As a first step to run the analysis, the geostatic stresses need to be
The above FE model was subjected to a seismic loading, which, for distributed in the wall–soil system. This has been performed by con-
this study, consisted of a real acceleration–time history of the 1989 sidering the construction sequence of a typical cantilever-type
Loma Prieta earthquake (PEER 2018), having a peak ground accel- retaining wall. To run the FE simulation, the input properties used
eration (PGA) of 0.264g [Fig. 2(b)] and dominant frequencies (f) of in the study were chosen as per Table 1. The wall was assumed to
about 0.7, 2.2, and 2.7 Hz [Fig. 2(c)]. To investigate the effects of be constructed in six stages, as presented in Fig. 3, in which the ini-
applied earthquake acceleration amplitude and frequency content tial stage relates to the placement of the foundation soil and installa-
(f) on the seismic response of the wall–soil system, scaled uniform tion of the wall. The subsequent four stages simulate the placement
sinusoidal acceleration–time histories were also considered with of the backfill soil in lifts of thickness 0.22 H for each stage,
three amplitudes of 0.2, 0.4, and 0.6g, and scaled frequencies of 0.5, whereas the last stage simulates the placement of the backfill soil in
2, and 4 Hz. These acceleration–time histories were applied at the a lift of thickness 0.11 H. The backfill soil layers was placed in the
base of the FE model, as presented in Fig. 2(a). The effect of vertical FE model and simulated by using the physical parameters measured
seismic acceleration was neglected in the present analysis. after the compaction process in the centrifuge test conducted by Jo
Notionally, the vertical seismic acceleration may contribute to the et al. (2014). Therefore, the compaction process was already taken

© ASCE 04019117-4 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


Table 1. Soil and wall parameters used for the present study

Parameter Symbol Unit Value


Soil
Dry unit weight gs kN/m3 14.23
Angle of shearing resistance f0 degrees 40.00
Dilatancy c degrees 10.00
Young’s modulus at 50% failure strength, corresponding to pref E50ref
MPa 46.80
ref
Odometeric modulus, corresponding to pref Eod MPa 46.80
Modulus for unloading–reloading conditions, corresponding to pref Eurref
MPa 140.40
Small strain shear modulus, corresponding to pref Gref
o MPa 113.00
Reference shear strain, corresponding to 70% of Gref
o g 0.7 — 0.0002
Poisson’s ratio for unloading–reloading conditions  ur–s — 0.20
Damping ratio zs % 3.00
Reference confining pressure pref kN/m2 100.00
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

Stiffness stress-level dependency constant K — 0.50


Failure ratio Rf — 0.90
Wall
Unit weight g wall kN/m3 26.60
Modulus of elasticity Ewall GPa 68.00
Second moment of area of the stem Istem m4 8.873  10–4
Poisson’s ratio  wall — 0.334
Damping ratio z wall % 3.00

Fig. 3. Construction sequence of a typical cantilever-type retaining wall–soil system and the associated contours of horizontal displacement.

into account during the static analysis. The soil behavior was also Stage 5, the movement of the stem away from the backfill soil
simulated by using the Hssmall model. This model can replicate the appears to be more than the backfill soil in the upper part (the nega-
volumetric mechanism (cap). The main role of the volumetric tive sign in Fig. 3 means that the displacement is away from the
mechanism (cap) is to close the elastic domain and simulate the den- backfill soil), perhaps because of the development of the earth pres-
sification/compaction of the material. With the progressive place- sure thereby causing an elastic deformation of the stem. However,
ment of the backfill soil, the wall will be subject to displacement in the lower parts of the stem, it is observed that the backfill soil
and rotation, as presented in Fig. 3. It can be noted from Fig. 3 that moves more than the stem of the wall. From Stages 4 and 5 of
during the placement of the backfill soil in lifts of 0.22-H thickness, Fig. 3, a clear formation of a failure plane originating from the heel
the maximum deformations are mobilized in the backfill soil above of the wall and extending up to the ground level at an inclination of
the base slab because the stem as well as the base slab of the wall an angle less than 45° to the horizontal is also observed. The defor-
rotate toward the backfill soil; this trend continues until Stage 4. In mation behavior of the wall matches with what has been observed

© ASCE 04019117-5 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


in the real field observations reported by Bentler and Labuz (2006). phenomenon was due to the interlocking grain fabric of the dry pluvi-
The deformation of the wall is also predicted at different durations of ated sand. The numerical modeling simulation of the present study,
earthquake and discussed later in the seismic analysis. After the simu- also showed similar results, but these are attributed to the complicated
lation of the construction stages and consequent distribution of the deformation mechanism during the construction process of a
geostatic stresses in the FE model, the static earth pressures are com- cantilever-type retaining wall. This is depicted very clearly in Fig. 3.
puted by the FE model, both behind the stem, pstem(static), and along the As can be observed from Fig. 3 for the last stage of construction pro-
virtual plane, pvp(static), which are then compared with the Rankine so- cess, the wall rotates as a rigid body about the heel toward the backfill
lution as well as with the Jo et al. (2014) centrifuge test results. The soil while the stem rotates about the base away from the backfill soil
earth pressure comparison is presented in Figs. 4(a and b) for walls of because of the elastic deflection. Therefore, the upper part of the stem
heights 5.4 m and 10.8 m, respectively. From Figs. 4(a) and 6(b), it is has a sufficient lateral movement away from the backfill soil, causing
observed that the FE model predictions are in an excellent agreement the development of active earth pressure. However, for the lower sec-
with the Jo et al. (2014) centrifuge test results. In addition, it is tion of the wall, the rotation of wall as rigid body about the heel to-
observed that pstem(static) and pvp(static) values obtained by the FE model ward the backfill soil is dominant, and the stem has insufficient lateral
in the top ¾ H are very close to the static active earth pressure values yielding away from the backfill compared with the upper part of the
obtained by Rankine’s theory, whereas for the bottom ¼ H, the earth stem causing that the lateral earth pressure to remain close to the at-
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

pressure values are between the static active value, also obtained by rest state.
Rankine’s theory, and the at-rest value. Jo et al. (2014) explained this

Seismic Analysis

After performing the aforementioned static analysis, a seismic anal-


ysis was carried out for the FE model, which, as mentioned above,
was carried out by applying a seismic loading in the form of an
acceleration–time history at the base of the FE model. Through the
seismic analysis, apart from obtaining Pstem, Pvp, Fwa, Fwp, Fsa, Fsp,
Nwall, and Mwall, the acceleration and sliding response of the wall–
soil system has also been obtained and its importance is discussed
in the sections below. It is to be highlighted that for both the static
and seismic analyses, a mesh-size sensitivity analysis was carried
out until a convergent solution was achieved. The total earth pres-
sure force was chosen to be the parameter for the sensitivity analy-
sis, and based on this, the final FE mesh comprised of 4,418 ele-
ments with 9,164 nodes.

Acceleration Response of the Cantilever-Type


Retaining Wall–Soil System
The acceleration response of the wall–soil system is presented in
Fig. 5 for the time duration of 3–7 s. This is the duration in which
the intensity of the applied earthquake acceleration was concen-
trated [Fig. 2(b)] and, hence, it was chosen for presenting the results
from the seismic analysis. From Fig. 5, it is observed that the

Fig. 4. Comparison of static earth pressure profiles: (a) for wall height, Fig. 5. Acceleration response of the cantilever-type retaining wall–
H = 5.4 m; and (b) for wall height, H = 10.8 m. soil system.

© ASCE 04019117-6 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


acceleration response for the top of the stem and top of the backfill Fig. 6(c), the amplitude of the acceleration response for the top of
soil match each other. This implies that at the top of the FE model, the stem is much higher than that for the backfill soil. Similarly,
the stem of the wall and backfill soil move together; that is, they are from Figs. 6(d and g), it is observed that for an applied acceleration
in phase with each other. It is also observed that the acceleration of with 0.5 Hz frequency, the amplitudes of acceleration for the top of
the top of the stem and backfill soil is higher than the acceleration at the stem and backfill soil have the same amplitude as the applied
the base of the wall, thereby implying a possible amplification of acceleration amplitudes of 0.4g and 0.6g, respectively. However,
acceleration toward the top of the FE model. The acceleration for an applied acceleration with a frequency content of 2 Hz, as pre-
response of the wall–soil system, when it is subject to a uniform si- sented in Figs. 6(e) and 8(h), the amplitude of acceleration for the
nusoidal acceleration–time history of different amplitudes and fre- top of the stem is higher than the acceleration for the top of the back-
quency contents, is presented in Fig. 6. From Fig. 6(a) it is observed fill soil. When the frequency content of the applied acceleration is
that when the amplitude of the applied earthquake acceleration is further increased to 4 Hz for applied acceleration amplitudes of 0.4g
0.2g with the frequency content of 0.5 Hz, the amplitude of the and 0.6g, as presented in Figs. 6(f) and 8(j), respectively, it is
acceleration response for both the top of the stem and backfill soil observed that the amplitude of acceleration for the stem amplifies to
matches the amplitude of the applied earthquake acceleration itself. a maximum value of 1.8g (i.e., it becomes more than the amplitude
However, when the frequency content of the applied earthquake of the applied acceleration). On the other hand, the amplitude of
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

acceleration is increased by four times to 2 Hz while the amplitude acceleration for the top of the backfill soil seems to deamplify, and
of the applied acceleration is kept same at 0.2g, the amplitude of the its maximum value becomes less than the amplitude of the applied
acceleration response for the top of the stem and backfill soil ampli- earthquake acceleration. This behavior—of acceleration amplifica-
fies to a value close to 0.4g, as presented in Fig. 6(b). On a further tion for the top and a deamplification for the bottom of the FEM—
increase of the frequency content to 4 Hz, with the amplitude of the reflects a nonlinear soil behavior that deamplifies a strong earth-
applied acceleration remaining the same at 0.2g, as presented in quake, resulting in a higher dissipation of the seismic energy. It is to

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Fig. 6. Acceleration response for top of the wall and backfill soil for the uniform sinusoidal acceleration–time history of different amplitudes and fre-
quency contents: (a–c) a = 0.2g; (d–f) a = 0.4g; and (g–i) a = 0.6g.

© ASCE 04019117-7 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


be noted that similar deamplification behavior of soil for strong soil that are in contact with the stem of the wall as well those that are
earthquakes has also been reported by Griffiths et al. (2016) and along the virtual vertical plane. These elemental lateral stresses are
Stamati et al. (2016). multiplied with the corresponding element heights, to get the ele-
mental seismic earth pressure forces. These elemental seismic earth
Wall and Backfill Seismic Inertia Forces: Fwa, Fwp, Fsa, pressure forces are summed together to get Pstem and Pvp, and their
and Fsp variation with time is presented in Figs. 8(a and b), respectively.
From Fig. 8(a), it is observed that at the beginning of the seismic
The wall and backfill seismic inertia forces—Fwa, Fwp, Fsa, and analysis (i.e., at time t = 0 s), Pstem is about 53 kN/m, which is
Fsp—are estimated by using the following procedure: first, elemen- between the static active and at-rest earth pressure force; at time t =
tal acceleration, awe and ase, is obtained by the FE model: for the 3.9 s, when the applied earthquake acceleration has a maximum
wall, awe is obtained for all the elements of the base slab and stem, value and is applied toward the backfill soil, Pstem has a maximum
whereas for the backfill soil, ase is obtained only for those elements value of 112 kN/m, whereas it attains a minimum value of about
that lie in the middle of backfill soil above the base slab. The corre- 45 kN/m when the applied earthquake acceleration has a maximum
sponding masses of the elements (for the case of the wall) and value but is applied away from the backfill soil at time t = 4.5 s. As
masses of the horizontal strips (for the case of the backfill soil) are discussed above, the cantilever-type retaining walls are designed by
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

multiplied with the elemental accelerations to get the elemental considering the same concepts as used for the design of rigid retain-
seismic inertia force. These elemental seismic inertia forces are ing walls. However, the aforementioned present study results—
summed together, both for the wall and backfill soil, to get the seis- which are in contrast with the observation of Nakamura (2006), who
mic inertia forces Fwa, Fwp (for the wall) and Fsa, Fsp (for the back- observed that for a rigid retaining wall, the maximum seismic earth
fill soil). As presented in Fig. 7, the wall and backfill seismic inertia pressure force is developed when the applied earthquake accelera-
forces are dependent upon the applied earthquake acceleration. It is tion is maximum but applied away from the backfill soil—show that
observed from Fig. 7 that the maximum wall and backfill seismic Pstem is maximum when the applied acceleration is applied toward
inertia forces are acting in an active direction (Fwa = 18.3 kN/m and the backfill soil. Thus, an active state is not developed behind the
Fsa = 54.2 kN/m) when the applied earthquake acceleration has a
maximum value and is applied toward the backfill soil at t = 3.9 s.
However, the maximum wall and backfill seismic inertia forces are
acting in a passive direction (Fwp = 36.7 kN/m and Fsp = 88.5 kN/m)
when the applied earthquake acceleration has a maximum value and
is applied away from the backfill soil at t = 4.5 s. It is observed from
Fig. 7 that the wall seismic inertia force (i.e., Fwa, Fwp) and backfill
seismic inertia force (i.e., Fsa, Fsp) are in phase, which implies that
the wall and backfill soil move as one entity. This is an extremely
important finding because this will significantly affect the develop-
ment of the active state in the wall–soil system when the earthquake
acceleration is applied toward the backfill soil, as discussed below.

Seismic Earth Pressure Forces: Pstem and Pvp


The seismic earth pressure force behind the stem, Pstem, and seismic
earth pressure force at the virtual plane, Pvp, have been estimated by
adopting the following procedure. First, the elemental lateral stresses
are obtained from the FE model for all those elements of the backfill (a)

(b)

Fig. 8. Seismic earth pressure forces: (a) at the stem, Pstem; and (b) at
Fig. 7. Wall and backfill seismic inertia forces: Fwa, Fwp, Fsa, and Fsp. the vertical plane, Pvp.

© ASCE 04019117-8 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


stem despite the fact that the acceleration is applied toward the back- virtual plane. On the other hand, at time t = 4.5 s, when the applied
fill soil, and, consequently, the retaining wall moves away from the earthquake acceleration has a maximum value and is applied away
backfill soil. The present study observations are also in contrast with from the backfill soil, a minimum load case is developed behind the
what was observed for the behavior of a cantilever-type retaining stem, whereas a maximum load case is developed at the vertical vir-
wall modeled via a numerical model by Green et al. (2008) and via tual plane. This clearly points to the fact that Pstem and Pvp attain
an experimental work by Kloukinas et al. (2015)—both of these maximum and minimum values at different times, thus suggesting
studies reported that a maximum value of Pstem is same to what is that there is a phase difference between these two quantities.
observed for a rigid retaining wall. Fig. 8(b) shows the variation of Therefore, for the purpose of structural and global stability, they
Pvp with time. It is observed that at the beginning of the seismic anal- have to be assessed individually.
ysis (i.e., at time t = 0 s), Pvp is about 60 kN/m, which, like Pstem, is It is also observed from Figs. 8(a and b), that at time t = 30 s, that
between the static active and at-rest state earth pressure force. At is, at the end of the seismic analysis, there is a residual Pstem and Pvp
time t = 3.9 s, when the applied earthquake acceleration has a maxi- of about 88 kN/m and 100 kN/m, respectively. These residual seismic
mum value and is applied toward the backfill soil, Pvp has a mini- earth pressure forces are attributed to the densification of backfill soil
mum value of 61 kN/m, whereas it attains a maximum value of about during the earthquake, thereby implying that the chosen HSsmall
165 kN/m when the applied earthquake acceleration has a maximum constitutive model caters to the simultaneous densification and settle-
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

value but is applied away from the backfill soil at time t = 4.5 s. ment during a seismic analysis. Figs. 9(a, c, and e) show the effect of
These observations are similar to the observations of a rigid retaining varying earthquake amplitude and frequency content on Pstem, and
wall as observed via a centrifuge test carried out by Nakamura the same for Pvp is presented in Figs. 9(b, d, and f). From Figs. 9(a–f),
(2006). Thus, it can be said that at time t = 3.9 s, when the applied it is observed that for all amplitudes and frequencies of the applied
earthquake acceleration has a maximum value and is applied toward earthquake, Pstem is maximum when the earthquake acceleration is
the backfill soil, a maximum load case is developed behind the stem applied toward the backfill soil, whereas Pstem is minimum when the
of the wall, whereas a minimum load case is developed at the vertical earthquake acceleration is applied away from the backfill soil. On the

(a) (b)

(c) (d)

(e) (f)

Fig. 9. Seismic earth pressure forces, Pstem and Pvp, for the uniform sinusoidal acceleration–time history of different amplitudes and frequency con-
tents: (a and b) f = 0.5 Hz; (c and d) f = 2 Hz; and (e and f) f = 4 Hz.

© ASCE 04019117-9 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


other hand, Pvp is maximum when the earthquake acceleration is DPstem and DPvp, defined as DPstem = Pstem – Pstem(static) and DPvp =
applied away from the backfill soil, and Pvp is minimum when the Pvp – Pvp(static), where Pstem(static) and Pvp(static) = static earth pres-
earthquake acceleration is applied toward the backfill soil. It is further sure force acting at the stem and vertical virtual plane, respectively.
observed that Pstem and Pvp increase significantly when the amplitude Figs. 10(a and b) and 10(c and d) show the variation of seismic earth
of the applied earthquake is increased from 0.2g to 0.4g, whereas on pressure force increments, DPstem and DPvp, and wall and backfill
further increasing the amplitude to 0.6g, Pstem and Pvp do not change seismic inertia forces, Fwa, Fwp Fsa, and Fsp, for the top 1=3Hstem and
with the same proportion as before—again indicating a possible bottom 1=3 Hstem, respectively. From Fig. 10(a and b) it is observed
deamplification of the acceleration response of the backfill soil for a that the seismic earth pressure force increment DPstem and wall seis-
strong earthquake. Both Pstem and Pvp are moderately sensitive to the mic inertia force, Fwa and Fwp, for the top 1=3 Hstem are out of phase
number of acceleration cycles; on the other hand, Pstem is highly sen- from each other, whereas for the bottom 1=3 Hstem, DPstem is in phase
sitive to the natural frequency of the wall. It is observed that Pstem with the wall seismic inertia forces. The reason for this disparity
increases with an increment in the frequency content of earthquake could be that the stem of the wall is monolithically fixed with the
acceleration and its maximum value is predicted when the earthquake base slab and thereby not allowing any relative displacement
acceleration is applied with a maximum frequency content and its between the stem and backfill soil. Similarly, from Figs. 10(c and d)
value is close to the natural frequency of the wall. These observations it is observed that the seismic earth pressure force increment DPvp
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

are in contrast to the observations reported by Cakir (2013), who, and soil seismic inertia forces, Fsa and Fsp, do not act in phase.
based on an FE study, found that the stresses behind the wall tend to
increase when the frequency content of earthquake acceleration Shear Force, Nwall, and Bending Moment, Mwall
decreases. However, Pvp appears to be not significantly affected by
the natural frequency of the wall. Figs. 11(a and b) respectively show the shear force Nwall and bending
moment Mwall time history predicted at the stem of the wall.
Studying Figs. 11(a and b) in conjunction with Fig. 8, it is observed
Phase Difference between the Seismic Earth Pressure
that the shear force Nwall and bending moment Mwall have the exact
Force Increments, DPstem and DPvp, and Wall and
same trend as the seismic earth pressure force Pstem. In addition, at
Backfill Seismic Inertia Forces, Fwa, Fwp, Fsa, and Fsp
time t = 3.9 s, when the applied earthquake acceleration has a maxi-
This section details the phase difference between various forces act- mum value and is applied toward the backfill soil, both Nwall
ing on the wall under seismic conditions. In order to clearly under- and Mwall attain their maximum values of about 120 kN/m and
stand the contribution of the seismic earth pressure force, it is dis- 220 kN·m/m, respectively, whereas they attain their minimum val-
cussed in terms of the seismic earth pressure force increments, ues of about 25 kN/m and 64 kN·m/m, respectively, at time t = 4.5 s,

(a) (b)

(c) (d)

Fig. 10. Phase difference between (a and b) DPstem, Fwa, and Fwp for the top and bottom 1=3 Hstem; and (c and d) DPvp, Fsa, and Fsp for the top and bottom 1=3 H.

© ASCE 04019117-10 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


is presented in Fig. 13. From the discussion in the preceding sec-
tions, it is clear that along the height of the stem, Hstem, there is an
amplification of acceleration toward the top of the stem—the same
is presented in Fig. 13. In addition, the variation of the seismic earth
pressure along the height of the stem may be considered to be linear,
as presented in Fig. 13. It is clear that Nwall and Mwall will vary along
the height of the stem and with the time of the earthquake. Further,
both of these quantities are dependent upon the wall seismic inertia
forces, Fwa and Fwp. For an accurate seismic structural stability anal-
ysis of the wall, it is crucial that the contribution of wall seismic iner-
tia forces should be evaluated at different locations along the height
of stem and at different times of the earthquake. Assuming that the
stem behaves as a cantilever beam, having a fixed connection with
the base slab, the shear force, Nw, and bending moment, Mw, time
history can be computed by using the dynamic Euler–Bernoulli
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

beam theory as shown by Eqs. (3) and (4), respectively:


(a) ∂3 ux ðz; tÞ
Nwall ðz; tÞ ¼ Ewall Istem
∂z3
ðt Hðstem  
∂2 ux ðz; tÞ
¼ m þ pstem ðz; tÞ dzdt (3)
∂t2
0 0

∂2 ux ðz; tÞ
Mwall ðz; tÞ ¼ Ewall Istem
∂z2
ðt Hðstem Hðstem  
∂2 ux ðz; tÞ
¼ m þ pstem ðz; tÞ dz2 dt (4)
∂t2
0 0 0

where Nwall(z, t) and Mwall(z, t) = shear force and bending moment


on the stem for time t and at height z along the stem; Ewall =
Young’s modulus of the wall; Istem = second moment of area of the
stem; ux(z, t) = horizontal elastic deflection of the stem for time t
and at height z along the stem; ∂2 ux ðz; tÞ=∂t2 = predicted horizontal
(b) acceleration for time t and at height z along the stem = astem ðz; tÞ;
astem ðz; tÞ = acceleration of the stem for time t and at height z along
Fig. 11. (a) Shear force, Nwall; and (b) bending moment, Mwall.
the stem; and pstem ðz; tÞ = seismic earth pressure for time t and at
height z along the stem. It is to be noted that z is measured above the
that is, when the applied earthquake acceleration has a maximum base slab along the height of the stem, Hstem. Eqs. (3) and (4) have
value and is applied away from the backfill soil. Thus, it can be been integrated numerically; and thus, if N is the number of ele-
argued that a critical case for the structural stability of a cantilever- ments of the stem, then Eqs. (3) and (4) could be rewritten
type retaining wall is when the maximum acceleration is applied to-
ward the backfill soil. Figs. 12(a, c, and e) and 12(b, d, and f) show X
N X
N

the effect of varying earthquake amplitude and frequency content on Nwall ðz; tÞ ¼ mn  astem n ðtÞ þ Pstem n ðtÞ (5)
1 1
Nwall and Mwall. It is observed that Nwall and Mwall show the same
trends as were observed for the Pstem and also that both Nwall and
Mwall are highly sensitive to the amplitude of the applied earthquake X
N X
N
when its value is between 0.2g and 0.4g. For an applied earthquake Mwall ðz; tÞ ¼ mn  astem n ðtÞ  zn þ Pstem n ðtÞ  zn
acceleration of amplitude > 0.4g, Nwall and Mwall do not remain as 1 1

sensitive as before—again concreting the fact that deamplification (6)


effects creep in for strong earthquakes. It can also be noted that Nwall
and Mwall, like Pstem and Pvp, are not sensitive to the number of where mn = mass of the nth element of the stem; astem n ðtÞ = accelera-
acceleration cycles where the maximum values of shear force and tion of the nth element of the stem for time t; zn = height of the nth ele-
bending moment are still having the same rate with increasing of ment along the stem; and Pstem_n(t) = seismic earth pressure force for
acceleration cycles. the nth element for time t. Eqs. (5) and (6) can be further simplified as

Effect of Wall Seismic Inertia Forces, Fwa and Fwp, and X


n

Seismic Earth Pressure Force, Pstem, on the Shear Force, Nwall ðz; tÞ ¼  g wall  bstem  ðzn  zn1 Þ  astem n ðtÞ
1
Nwall, and Bending Moment, Mwall
X
n
A free-body diagram of the stem of a cantilever-type retaining wall– þ Pstem n ðtÞ (7)
soil system showing various forces acting on it during an earthquake 1

© ASCE 04019117-11 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


(a) (b)
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

(c) (d)

(e) (f)

Fig. 12. Shear force, Nwall, and bending moment, Mwall, for the uniform sinusoidal acceleration–time history of different amplitudes and frequency
contents: (a and b) f = 0.5 Hz; (c and d) f = 2 Hz; and (e and f) f = 4 Hz.

X
n
force increment at the stem, DPstem [=Pstem – Pstem(static)] on shear
Mwall ðz; tÞ ¼  g wall  bstem  ðzn  zn1 Þ  astem n ðtÞ  zn force Nwall and bending moment Mwall of the stem. As observed,
1
at the beginning of the seismic analysis (i.e., time t = 0 s), only
X
n Pstem(static) causes Nwall and Mwall because other values are 0;
þ Pstem n ðtÞ  zn (8) however, at time t = 3.9 s, when the applied earthquake accelera-
1
tion has a maximum value and is applied toward the backfill soil,
Pstem(static), Fstem_a, and DPstem all contribute to Nwall and Mwall. In
where g wall = unit weight of the wall; and bstem = width of the stem.
addition, these Nwall and Mwall for time t = 3.9 s act away from the
From Eqs. (7) and (8), it is clear that the shear force and bending
backfill soil (i.e., in a direction opposite to the direction of the
moment depend upon the wall seismic inertia force and the seismic
applied earthquake acceleration and they have negative values, as
earth pressure force. Their effects are discussed next for the top and
shown in Table 2). When the applied earthquake acceleration has
bottom 1=3Hstem as well as for the mid-height of the stem of the wall.
a maximum value but is applied away from the backfill soil at
time t = 4.5 s, then, like before, Pstem(static), Fstem_p, and DPstem all
Effect of Stem Seismic Inertia Force, Fstem_a and Fstem_p, contribute to Nwall and Mwall. However, unlike the previous case,
on Shear Force, Nwall, and Bending Moment, Mwall, for Pstem(static) and DPstem produce Nwall and Mwall in the same direc-
the Top 1=3 Hstem tion as the direction of the applied earthquake acceleration and
Table 2 shows the relative contributions of static earth pressure they have negative values, as shown in Table 2, whereas Fstem_p
force for the stem Pstem(static), seismic stem inertia force Fstem_a, produces Nwall and Mwall in a direction opposite to the direction of
Fstem_p, where Fstem_a, Fstem_p are the stem seismic inertia force act- the applied earthquake acceleration and it has a positive value, as
ing away and toward the backfill soil, and seismic earth pressure shown in Table 2.

© ASCE 04019117-12 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


Effect of Stem Seismic Inertia Force, Fstem_a and Fstem_p, shown in Table 2). When the applied earthquake acceleration has a
on Shear Force, Nwall, and Bending Moment, Mwall, for maximum value but is applied away from the backfill soil at time t =
the Mid-Height of the Stem 4.5 s, like at the top 1=3 Hstem, Pstem(static), Fstem_p, and DPstem all con-
Table 2 shows the relative contributions of static earth pressure tribute to Nwall and Mwall. However, Pstem(static) and DPstem produce
force for the stem Pstem(static), stem seismic inertia force Fstem_a, Nwall and Mwall in the same direction as the direction of the applied
Fstem_p, and seismic earth pressure force increment DPstem on Nwall earthquake acceleration and they have negative values, as shown in
and Mwall for the mid-height of the stem. At the beginning of the Table 2, whereas Fstem_p produces Nwall and Mwall in a direction op-
seismic analysis (i.e., time t = 0 s), only Pstem(static) causes Nwall and posite to the direction of the applied earthquake acceleration and it
Mwall. However, at time t = 3.9 s, when the applied earthquake has a positive value, as shown in Table 2.
acceleration has a maximum value and is applied toward the backfill
soil, Pstem(static), Fstem_a, Fstem_p, and DPstem all contribute to Nwall Effect of Stem Seismic Inertia Force, Fstem_a and Fstem_p,
and Mwall, but the effect of Fstem_a on Nwall and Mwall is reduced at on Shear Force, Nwall, and Bending Moment, Mwall, for
the top 1=3 Hstem. In addition, all of these quantities act away from the Bottom 1=3 Hstem
the backfill soil (i.e., in a direction opposite to the direction of the Table 2 shows the effect of static earth pressure force Pstem(static),
applied earthquake acceleration and they have negative values, as stem seismic inertia force Fstem_a, Fstem_p, and seismic earth pres-
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

sure force increment DPstem on the Nwall and Mwall for the bottom
1=3 H
stem. At the beginning of the seismic analysis (i.e., time t = 0 s)
for the top 1=3 Hstem and mid-height of stem, only Pstem(static) causes
Nwall and Mwall; however, at time t = 3.9 s, when the applied earth-
quake acceleration has a maximum value and is applied toward the
backfill soil, the contribution of Fstem_a to Nwall and Mwall is very
small compared with the effect of Pstem(static) and DPstem. In addi-
tion, all of these quantities act away from the backfill soil and they
have negative values, as shown in Table 2. When the applied earth-
quake acceleration has a maximum value but is applied away from
the backfill soil at time t = 4.5 s, observations for the bottom
1=3 H
stem are similar to the top =3 Hstem and mid-height of the stem.
1
From the above, it can be concluded that when the earthquake
acceleration is applied toward the backfill soil, then for the top half
of the wall, the wall seismic inertia force (Fstem_a or Fstem_p) has a
major contribution to the shear force Nwall and bending moment
Mwall, whereas for the bottom half of the wall it is the combination
of static earth pressure force Pstem(static) and seismic earth pressure
force increment DPstem that contribute to the shear force, Nwall, and
Fig. 13. Free-body diagram of the stem of a cantilever-type retain- bending moment, Mwall. When the earthquake acceleration is
ing wall–soil system showing various forces acting on it during an applied away from the backfill soil, the stem seismic inertia force
earthquake. produces shear force, Nwall, and bending moment, Mwall, in the
direction of the static earth pressure force and the increment of

Table 2. Effect of stem seismic inertia force, fstem_a and fstem_p, on shear force, nwall, and bending moment, mwall, for the top 1=3Hstem, the mid-height of the
stem, and the bottom 1=3Hstem

Contribution to Nw (kN/m) due to Contribution to Mw (kN·m/m) due to


Time [t (s)] Nwall (kN/m) Pstem(static) Fstem_a or Fstem_p DPstem Mwall (kN·m/m) Pstem(static) Fstem_a or Fstem_p DPstem
1=3
Top Hstem
0.0 −0.72 −0.72 0.01 0.00 −3.68 −3.68 0.02 0.00
3.9 −2.65 −0.72 −1.08 −0.85 −14.50 −3.68 −8.89 −1.93
4.5 −0.88 −0.72 3.72 −3.88 −7.68 −3.68 17.22 −21.22
30.0 −2.20 −0.72 0.00 −1.48 −13.11 −3.68 0.00 −9.43
Mid-height of the stem
0.0 −9.45 −9.45 0.02 0.00 −36.32 −36.40 0.08 0.00
3.9 −26.79 −9.45 −4.68 −12.68 −110.68 −36.40 −24.79 −49.49
4.5 −10.51 −9.45 11.33 −12.39 −42.61 −36.40 47.94 −54.15
30.0 −22.70 −9.45 0.00 −13.52 −88.95 −36.40 0.00 −52.55
Bottom 1=3 Hstem
0.0 −52.97 −53.10 0.04 0.00 −92.34 −92.45 0.11 0.00
3.9 −122.90 −53.10 −11.86 −58.20 −217.60 −92.45 −29.34 −95.78
4.5 −28.17 −53.10 21.75 3.18 −63.90 −92.45 64.53 −35.98
30.0 −88.56 −53.10 0.00 −53.46 −173.90 −92.45 0.00 −81.45
Note: For all forces, a negative value indicates that the force is acting away from the backfill soil, whereas a positive value indicates that the force is acting
toward the backfill soil. Similarly, for the moments, a negative value indicates that the moment has a counterclockwise-sense, whereas a positive value indi-
cates that the moment has a clockwise-sense.

© ASCE 04019117-13 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


seismic earth pressure force, causing the shear force and bending foundation (Dw–f). As the amplitude of the applied earthquake accel-
moment to attain a minimum value of less than the static value. eration increases from 0.2g to 0.6g, Dw–f increases, whereas with an
increase in the frequency content of the applied earthquake accelera-
tion from 0.5 Hz to 4 Hz, the Dw–f decreases. This is in contrast to
Relative Displacement of the Wall and Backfill Soil with what has been observed for the shear force Nwall and bending
respect to the Foundation Soil, Dw–f and Ds–f moment Mwall, which, as described in the preceding sections, attain
For the wall and backfill soil, relative displacement profiles were maximum values when both the frequency content and amplitude of
constructed for the following pairs: the applied earthquake are maximum (Fig. 12). It is also interesting
• The base of the wall and a reference point 0.5 m below the to note that Dw–f is about 0.2 m for an applied earthquake amplitude
base of the wall in the foundation soil [i.e., between base_wall of 0.6g and a frequency content of 4 Hz [Fig. 15(c)], whereas the
and P1 in Fig. 1(a)]. same for an applied earthquake amplitude of 0.4g and a frequency
• The center of gravity of the backfill soil and a reference point content of 2 Hz [Fig. 15(b)] is about 0.25 m, thereby suggesting that
0.5 m below the base of the wall in the foundation soil [i.e., the frequency content of the applied earthquake is a more dominating
between s_CG and P2 in Fig. 1(a)]. factor than its amplitude that contributes to Dw–f. From the above
These relative displacement profiles for the above pairs are pre- observations, it can be argued that for the global stability of a
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

sented in Figs. 14(a and b). It is observed from Fig. 14(a) that a maxi- cantilever-type retaining wall a low-frequency content of the applied
mum relative displacement of about 0.035 m between the wall and earthquake creates a critical case scenario, whereas, for the structural
foundation (Dw–f) occurs at time t = 3.9 s, which is the same time at stability of the cantilever-type retaining wall a high-frequency con-
which Pvp is minimum [Fig. 8(b)]. In addition, the relative displace- tent of applied earthquake creates a critical case scenario. The results
ment between the backfill and foundation soil (Ds–f) achieves its also show that the sliding of the wall–soil system is highly sensitive
maximum value of about 0.025 m for time t = 3.9 s, and remains con- to the number of acceleration cycles (and the duration of the applied
stant until the end of the seismic analysis. Thus, from the above two earthquake acceleration), which is in contrast to what has been
observations, it can be said that the wall and backfill soil move as a
single entity and Pvp is causing the sliding of the wall–soil system.
Fig. 15 shows the effect of varying earthquake amplitude and fre-
quency content on the relative displacement between the wall and

(a)

(a)

(b)

(c)
(b)
Fig. 15. Wall–foundation relative displacement, Dw–f, for the uniform
Fig. 14. (a) Wall–foundation relative displacement, Dw–f; and (b) sinusoidal acceleration–time history of different amplitudes and fre-
soil–foundation relative displacement Ds–f. quency contents: (a) f = 0.5 Hz; (b) f = 2 Hz; and (c) f = 4 Hz.

© ASCE 04019117-14 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


observed for the structural stability where the shear force Nwall and together. It can be observed from Fig. 16 that the deformation mech-
bending moment Mwall are not sensitive to the acceleration cycles. anism of the cantilever-type retaining wall is quite complicated and
includes the sliding of the cantilever-type retaining wall relative to
Deformation Shapes of the Cantilever-Type Retaining the foundation layer, rotation of cantilever-type retaining wall as a
Wall–Soil System at Various Times during the Duration rigid body because of the foundation soil deformability, and the
of the Earthquake rotation of the stem due to its elastic deflection. At time t = 3.9 s,
when the earthquake acceleration is applied toward the backfill soil
Fig. 16 presents the deformation and contours of horizontal displace- and has a maximum value [Fig. 2(b)], the cantilever-type retaining
ment of the cantilever-type retaining wall–soil system at times t = wall slides away from the backfill soil to a maximum distance, the
3.9, 4.5, and 30 s (i.e., at the end of the seismic analysis). Because stem deflects by a maximum value away from the backfill soil, and
the dynamic analysis was carried out after the completion of all the the retaining wall rotates about the toe by a maximum value away
five stages of construction of the retaining wall, the deformation and from the backfill soil [Fig. 16(a)]. At the same time, the seismic
contours of horizontal displacement of the cantilever-type retaining earth pressure force increments behind the upper half of the stem
wall–soil system presented in Stage 5 of Fig. 3 can also be said to be DPstem and along the virtual vertical plane DPvp reach a minimum
the results for t = 0 s of the dynamic analysis. For Figs. 16(a–c), it is value and they are close to the static earth pressure [Figs. 10(a, c,
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

important to highlight that the deformation shape of the stem and


and d)]. Thus, it can be said that the seismic earth pressure forces in
base slab shown is measured relative to the original position of the
these locations reach the active states and their values do not change
retaining wall, that is, at the start of the dynamic analysis at t = 0 s.
with increasing acceleration level and wall movement. However, at
Stage 5 of Fig. 3 shows that at t = 0 s, the stem rotates by 0.02° away
the same time, t = 3.9 s, it can be observed that the seismic earth
from the backfill soil while the base slab heel has a clockwise rota-
pressure force increment, DPstem, in the lower half of the stem
tion of 0.065°—thereby suggesting that the stem and base slab rotate
reaches the maximum value and does not reach the active state
in opposite directions to each other). However, at time t = 3.9 s—
[Fig. 10(b)] and its value increases with increasing earthquake
Fig. 16(a)—when the earthquake acceleration has its maximum
acceleration and wall movement. Similarly, at time t = 4.5 s, when
value and is applied toward the backfill soil, the stem rotation is pre-
the earthquake acceleration is applied away from the backfill soil
dicted as 0.217° away from the backfill soil while the base slab toe
and it has a maximum value, the cantilever-type retaining wall
has a counterclockwise rotation of 0.014°—thereby suggesting that
between t = 0 and 3.9 s, both the stem and base slab rotate in the slides by a maximum value toward the backfill soil, the stem
same direction. In addition, as described in the previous section, at deflects by a maximum value toward the backfill soil, and the retain-
time t = 3.9 s, the wall slides as a rigid body away from the backfill ing wall rotates about the toe by a maximum value toward the back-
soil by about 0.025 m. At time t = 4.5 s, as presented in Fig. 16(b), fill soil. At the same time, the seismic earth pressures force incre-
when the earthquake acceleration has its maximum value and is ment, DPstem, in the upper half of the stem and along the virtual
applied away from the backfill soil, the stem rotates back toward the vertical plane reach a maximum value, and they partially mobilize
backfill soil (but is still away from its original position) by about the passive state [Figs. 10(a, c, and d)]. Thus, it can be said that the
0.017° while the base slab toe rotates clockwise (but still has a rota- seismic earth pressure in these locations partially mobilize the pas-
tion of about 0.01° compared with its original position at t = 0 s). sive state and their value increases with increasing earthquake
The wall slides toward the backfill soil; however, it is still away acceleration and wall movement. However, at the same time, t =
from the backfill soil by about 0.017 m as compared with its original 4.5 s, it can be observed that the seismic earth pressure force incre-
position. At the end of the seismic analysis at time t = 30 s, the stem ment, DPstem, in the lower half of the stem reaches the minimum
has a residual rotation of 0.204° relative to its original position and value [Fig. 10(b)] and its value remains close to the static earth pres-
the base slab toe has a residual counterclockwise rotation of 0.038° sure and does not change with increasing earthquake acceleration
in to the foundation soil while the wall has a residual sliding away and wall movement.
from the backfill soil of about 0.035 m, as presented in Fig. 16(c).
Effect of Height of the Cantilever-Type Retaining
Wall–Soil System, H, and the Natural Frequency, fn,
Relationship between the Seismic Earth Pressure and on Bending Moment, Mwall, and Wall–Foundation
Wall Displacement
Relative Displacement, Dw–f
In order to understand the relationship between the seismic earth All of the aforementioned analysis has been carried out for a wall of
pressure and wall displacement, Figs. 10 and 16 should be studied height H = 5.4 m. In order to assess the influence of the height and

Fig. 16. Deformation shapes of the cantilever-type retaining wall–soil system at various times during the duration of the earthquake: (a) t = 3.9 s;
(c) t = 4.5 s; and (d) t = 30 s.

© ASCE 04019117-15 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


(a) (b) (c)
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

(d) (e) (f)

Fig. 17. Bending moment, Mwall, for a uniform sinusoidal acceleration amplitude of 0.4g and different frequency content (a–c) for wall height, H =
5.4 m; and (d–f) for wall height, H = 10.8 m.

thereby the natural frequency of the wall on the structural and global • The study successfully identifies critical scenarios for the
stability under seismic loading, a wall of a different height of H = structural and global stability of a cantilever-type retaining
10.8 m is also analyzed using the aforementioned FEM, and the wall–soil system by considering earthquake characteristics,
comparison of results is presented herewith. From Figs. 17(c and e) analyzing the effects of seismic earth pressure and natural fre-
it is observed that for walls of height H = 5.4 m and 10.8 m, the bend- quency, and capturing the deformation mechanism.
ing moment, Mwall, is maximum (330 kN·m/m for H = 5.4 m and • For the seismic structural and global stability analyses of a
1,750 kN·m/m for H = 10.8 m, respectively) when the applied earth- cantilever-type retaining wall, the FEM has been innovatively
quake acceleration has a frequency content of 4 Hz (for H = 5.4 m) used, in which the structural and global stability of a
and 2 Hz (for H = 10.8 m), which are the frequencies close to the nat- cantilever-type retaining wall is analyzed by considering the
ural frequency of the respective wall–soil systems. However, from seismic earth pressure, computed at the stem (Pstem) and along
Figs. 18(a and d), which show the relative displacement between the a virtual plane (Pvp), and wall and backfill seismic inertia
wall and foundation soil, Dw–f, it is observed that for the wall of forces. It is noted that Pstem contributes to the structural stabil-
heights H = 5.4 m and 10.8 m, the maximum Dw–f is predicted when ity, whereas Pvp contributes to the global stability. It is also
the applied earthquake acceleration has a frequency content of observed that Pstem and Pvp are out of phase during the entire
0.5 Hz. In addition, for both H = 5.4 m and H = 10.8 m, it is observed duration of the earthquake.
that with increasing frequency content of the applied earthquake • A critical case for the structural stability occurs when the
acceleration, the relative displacement between the wall and founda- earthquake acceleration is directed toward the backfill soil and
tion soil, Dw–f, is decreased. From the above results, it is found that its frequency content is close to the natural frequency of the
the structural stability is affected by the natural frequency of the retaining wall. In contrast, for the global stability, the critical
wall, and a critical scenario will be when the natural frequency of the case occurs when the earthquake acceleration is maximum
wall is equal to the frequency of the applied earthquake acceleration. applied toward the backfill soil having the smallest frequency
On the other hand, it can be safely argued that the height of the wall content.
• The structural stability of the cantilever-type retaining wall is
does not affect the nature of the results for the global stability and
that the critical case will always occur when the frequency content found to be highly dependent on the natural frequency of the
of the applied earthquake acceleration has a minimum value. cantilever-type retaining wall relative to the applied earth-
quake frequency content, whereas the global stability does not
appear to be affected by it.
Conclusions and Recommendations • For the critical structural stability case, it has been observed
that the wall seismic inertia force has a significant contribution
• This paper presents a unique FE-based numerical modeling to the shear force and bending moment in the top half of the
approach to study both the seismic structural and global stabil- height of the stem. For the lower half of the height of the stem,
ity of a cantilever-type retaining wall–soil system. the seismic earth pressure contributed significantly to the shear

© ASCE 04019117-16 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


(a) (b) (c)
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

(d) (e) (f)

Fig. 18. Wall–foundation relative displacement, Dw–f, for a uniform sinusoidal acceleration amplitude of 0.4g and different frequency content
(a–c) for wall height, H = 5.4 m; and (d–f) for wall height, H = 10.8 m.

force and bending moment, whereas the contribution coming seismic earth pressure is not significant, and thus only static
from the wall seismic inertia force was very nominal. earth pressure, wall seismic inertia force, and backfill seismic
• When the earthquake acceleration is applied away from the inertia force should be considered as the total driving force
backfill soil and has its maximum value, the wall seismic iner- causing sliding instability to the wall–soil system.
tia force acts in a direction opposite of the seismic earth pres- • The effect of site characteristics should be considered during
sure, thereby causing a reduction of the shear force and the analysis of the seismic structural and global stability of a
bending moment. cantilever-type retaining wall (i.e., the amplification of accel-
• The number of acceleration cycles of the applied earthquake eration response of backfill soil at low acceleration level and
acceleration moderately affects the seismic earth pressure de-amplification of acceleration response of backfill soil at
behind the stem as well as the shear force and bending high acceleration level).
moment, whereas the relative displacement between the wall
and soil is observed to be highly sensitive to this. It is also
observed that the shear force and bending moment profiles Acknowledgments
match with the profiles of the seismic earth pressure behind
the stem. The first author would like to thank the Ministry of Higher
Based on the aforementioned conclusions, the following recom- Education and Scientific Research in Iraq for supporting their
mendations can be made: studies and funding this research. All the authors also thank the
• The seismic structural and global stability of a cantilever-type School of Mechanical, Aerospace and Civil Engineering, the
retaining wall should be checked separately, in which, the University of Manchester, for providing facilities to conduct the
structural stability should be checked by considering the maxi- research. The critique of the two anonymous reviewers is gratefully
mum earthquake acceleration anticipated at the construction acknowledged by the authors—this helped greatly in revising and
site and the frequency content of earthquake acceleration being improving the manuscript.
equal to the natural frequency of the structure, whereas for the
global stability, a check should be made by considering maxi- Notation
mum anticipated earthquake acceleration with a minimum fre-
quency content. The following symbols are used in this paper:
• The stem seismic inertia force should be considered for the a ¼ acceleration (g);
structural design of the upper half of the stem, whereas for an ¼ acceleration of the nth element (g);
lower half part, it could be neglected—thus proposing an eco- awe ¼ elemental acceleration for the elements of the
nomic yet safe design. base slab and stem (g);
• The seismic earth pressure is crucial for the structural stability ase ¼ elemental acceleration for the soil elements
especially for the lower half of the stem and should be consid- lying above and in the middle of the base
ered in the structural design. However, for global stability, slab (g);

© ASCE 04019117-17 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


astem_n(t) ¼ acceleration of the nth element of the stem for Ww ¼ weight of the wall (kN);
time t (g); ux(z, t) ¼ horizontal elastic deflection of the stem for
astem(z, t) ¼ acceleration of the stem for time t and at depth time t and at height z along the stem (m);
z along the stem (g); z ¼ height of the stem above the base slab (m);
bslab ¼ width of the base slab (m); zn ¼ height of the nth element along the stem (m);
bstem ¼ width of the stem (m); a, b ¼ Rayleigh damping parameters;
[C] ¼ damping FE matrix of the system; g s ¼ unit weight of the soil (kN/m3);
Dr ¼ relative density of the soil (%); g wall ¼ unit weight of the wall (kN/m3);
Ewall ¼ Young’s modulus of the wall (kN/m2); g 0.7 ¼ reference shear strain, corresponding to 70%
ref
E50 ¼ Young’s modulus at 50% failure strength of of Gref
o ;
soil, corresponding to pref (kN/m2); DPstem ¼ seismic earth pressure force increment at the
stem (kN/m);
Eod ¼ odometric modulus, corresponding to pref
ref
DPvp ¼ seismic earth pressure force increment at the
(kN/m2);
virtual plane (kN/m);
ref
Eur ¼ modulus for unloading–reloading conditions, Ds–f ¼ relative displacement between backfill and
corresponding to pref (kN/m2);
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

foundation soil (m);


Fsa ¼ backfill seismic inertia force acting toward the Dw–f ¼ relative displacement between wall and foun-
wall (kN/m); dation soil (m);
Fsp ¼ backfill seismic inertia force acting away from z s, z wall ¼ damping ratio for the soil and wall, respec-
the wall (kN/m);
tively (%);
Fstem_a ¼ stem seismic inertia force acting away from
 wall ¼ Poisson’s ratio of the wall;
the backfill soil (kN/m);  ur–s ¼ Poisson’s ratio for unloading–reloading condi-
Fstem_p ¼ stem seismic inertia force acting toward the
tions of the soil;
backfill soil (kN/m);
w 0 ¼ angle of shearing resistance of the soil (°);
Fwa ¼ wall seismic inertia force acting away from
c ¼ dilatancy of the soil (°); and
the backfill soil (kN/m);
v z1, v z2 ¼ first two natural circular frequencies of the FE
Fwp ¼ wall seismic inertia force acting toward the
model (rad/s).
backfill soil (kN/m);
f ¼ frequency of applied earthquake acceleration
(Hz); References
o ¼ small strain shear modulus, corresponding to
Gref
pref (kN/m2); ABAQUS. 2013. Analysis user’s manual. Version 6.13. Dassault
g ¼ acceleration due to gravity (m/s2); Systemes.
Hstem ¼ height of the stem (m); Ahmad, S. M., and D. Choudhury. 2010. “Seismic rotational stability
of waterfront retaining wall using pseudodynamic method.” Int.
Istem ¼ second moment of area of the stem (m4);
J. Geomech. 10 (1): 45–52. https://doi.org/10.1061/(ASCE)1532
[K] ¼ stiffness FE matrix of the system; -3641(2010)10:1(45).
k ¼ stiffness stress-level dependency constant; Al Atik, L., and N. Sitar. 2009. Experimental and analytical study of the
[M] ¼ mass FE matrix of the system; seismic performance of retaining structures. PEER Rep. 2008/104,
Mwall ¼ bending moment on the stem (kN·m/m); Berkeley, California: Pacific Earthquake Engineering Research Center,
Mwall(z, t) ¼ shear force on the stem for time t and at height College of Engineering, Univ. of California, Berkeley.
z along the stem (kN·m/m); Bakr, J., and S. M. Ahmad. 2018a. “A finite element performance-based
mn ¼ mass of the nth element of the stem (ton); approach to correlate movement of a rigid retaining wall with seismic
Nwall ¼ shear force on the stem (kN/m); earth pressure.” Soil Dyn. Earthquake Eng. 114 (Nov): 460–479. https://
Nwall(z, t) ¼ shear force on the stem for time t and at height doi.org/10.1016/j.soildyn.2018.07.025.
z along the stem (kN/m); Bakr, J., and S. M. Ahmad. 2018b. “Effect of earthquake characteristics on
pref ¼ reference confining pressure (kN/m2); the permanent displacement of a cantilever retaining wall.” In Vol 1 of
Proc., Numerical Methods in Geotechnical Engineering IX, edited by
Pstem ¼ seismic earth pressure force at the stem (kN/m);
A. S. Cardoso, J. L. Borges, P. A. Costa, A. T. Gomes, J. C. Marques,
Pstem_n(t) ¼ seismic earth pressure force for the nth and C. S. Vierira, 849–854. Leiden, Netherlands: CRC Press.
element for time t (kN/m); Basha, B. M., and G. S. Babu. 2010. “Seismic rotational displacements of
Pstem(static) ¼ static earth pressure force at the stem (kN/m); gravity walls by pseudodynamic method with curved rupture surface.”
Pvp ¼ seismic earth pressure force at the virtual Int. J. Geomech. 10 (3): 93–105. https://doi.org/10.1061/(ASCE)GM
plane (kN/m); .1943-5622.0000037.
Pvp(static) ¼ static earth pressure force at the virtual plane Bentler, J. G., and J. F. Labuz. 2006. “Performance of a cantilever retaining
(kN/m); wall.” J. Geotech. Geoenviron. Eng. 32 (8): 1062–1070. https://doi.org
pstem(static) ¼ static earth pressure at the stem /10.1061/(ASCE)1090-0241(2006)132:8(1062).
(kN/m2); Brinkgreve, R. B. J., S. Kumarswamy, and W. Swolfs. 2016. Plaxis 2016.
pstem(z, t) ¼ seismic earth pressure for time t and at height Delft, Netherlands: Plaxis.
Cakir, T. 2013. “Evaluation of the effect of earthquake frequency content
z along the stem (kN/m2);
on seismic behavior of cantilever retaining wall including soil–structure
pvp(static) ¼ seismic earth pressure acting at the virtual interaction.” Soil Dyn. Earthquake Eng. 45 (Feb): 96–111. https://doi
plane (kN/m2); .org/10.1016/j.soildyn.2012.11.008.
Rf ¼ failure ratio; Choudhury, D., and A. D. Katdare. 2013. “New approach to determine seis-
Rinter ¼ interface strength-reduction factor; mic passive resistance on retaining walls considering seismic waves.”
Ws ¼ weight of the backfill soil above the base slab Int. J. Geomech. 13 (6): 852–860. https://doi.org/10.1061/(ASCE)GM
(kN); .1943-5622.0000285.

© ASCE 04019117-18 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117


Choudhury, D., and S. S. Nimbalkar. 2008. “Seismic rotational displace- Mononobe, N., and M. Matsuo. 1929. “On the determination of earth pres-
ment of gravity walls by pseudodynamic method.” Int. J. Geomech. sures during earthquakes.” In Vol. 9 of Proc., World Engineering Conf.,
8 (3): 169–175. https://doi.org/10.1061/(ASCE)1532-3641(2008)8: 177–185. Tokyo: Kogakkai.
3(169). Nadim, F., and R. V. Whitman. 1983. “Seismically induced movement of
Coulomb, C. A. 1776. “Essai sur une application des règles des maximis et retaining walls.” J. Geotech. Eng. 109 (7): 915–931. https://doi.org/10
minimis à quelques problèmes de statique, relatits à l’architecture.” In .1061/(ASCE)0733-9410(1983)109:7(915).
Memoires de Mathematique et de Physique, Presentees a l’Academie Nakamura, S. 2006. “Reexamination of Mononobe-Okabe theory of gravity
Royale des Sciences, par divers Savans, 343–382. [In French.] Paris: De retaining walls using centrifuge model tests.” Soils Found. 46 (2): 135–
l’Imprimerie Royale. 146. https://doi.org/10.3208/sandf.46.135.
Dey, A. K., A. Dey, and S. Sukladas. 2017. “3N Formulation of the horizon- Ni, P., G. Mei, and Y. Zhao. 2017. “Displacement-dependent earth pres-
tal slice method in evaluating pseudostatic method for analysis of seis- sures on rigid retaining walls with compressible geofoam inclusions:
mic active earth pressure.” Int. J. Geomech. 17 (1): 04016037. https:// Physical modeling and analytical solutions.” Int. J. Geomech. 17 (6):
doi.org/10.1061/(ASCE)GM.1943-5622.0000662. 04016132. https://doi.org/10.1061/(ASCE)GM.1943-5622.0000838.
Green, R. A., C. G. Olgun, and W. I. Cameron. 2008. “Response and model- Paruvakat, N., X. Zeng, and R. S. Steedman. 2001. “Rotating block
ling of cantilever retaining walls subjected to seismic motions.” method for seismic displacement of gravity walls.” J. Geotech.
Comput.-Aided Civil Infrastruct. Eng. 23 (4): 309–322. https://doi.org Geoenviron. Eng. 127 (11): 994–995. https://doi.org/10.1061
Downloaded from ascelibrary.org by 91.106.47.101 on 10/03/22. Copyright ASCE. For personal use only; all rights reserved.

/10.1111/j.1467-8667.2007.00538.x. /(ASCE)1090-0241(2001)127:11(994).
Griffiths, S. C., B. R. Cox, and E. M. Rathje. 2016. “Challenges associated PEER (Pacific Earthquake Engineering Research Center). 2018. “PEER
with site response analyses for soft soils subjected to high-intensity input Ground Motion Database.” Accessed October 20, 2018. https://
ground motions.” Soil Dyn. Earthquake Eng. 85 (Jun): 1–10. https://doi ngawest2.berkeley.edu.
.org/10.1016/j.soildyn.2016.03.008. Rajasekaran, S. 2009. Structural dynamics of earthquake engineering:
Huang, C.-C. 2005. “Seismic displacements of soil retaining walls situated Theory and application using Mathematica and Matlab. Oxford, UK:
on slope.” J. Geotech. Geoenviron. Eng. 131 (9): 1108–1117. https://doi Woodhead Publishing.
.org/10.1061/(ASCE)1090-0241(2005)131:9(1108). Rajesh, B. G., and D. Choudhury. 2017. “Generalized seismic active thrust
Huang, C.-C., S.-H. Wu, and H.-J. Wu. 2009. “Seismic displacement crite- on a retaining wall with submerged backfill using a modified pseudody-
rion for soil retaining walls based on soil strength mobilization.” J. namic method.” Int. J. Geomech. 17 (3): 06016023. https://doi.org/10
Geotech. Geoenviron. Eng. 135 (1): 74–83. https://doi.org/10.1061 .1061/(ASCE)GM.1943-5622.0000750.
/(ASCE)1090-0241(2009)135:1(74). Stamati, O., N. Klimis, and T. Lazaridis. 2016. “Evidence of complex site
Jo, S.-B., J.-G. Ha, J.-S. Lee, and D.-S. Kim. 2017. “Evaluation of the seis- effects and soil non-linearity numerically estimated by 2D vs 1D seismic
mic earth pressure for inverted T-shape stiff retaining wall in cohesion- response analyses in the city of Xanthi.” Soil Dyn. Earthquake Eng. 87
less soils via dynamic centrifuge.” Soil Dyn. Earthquake Eng. 92 (Jan): (Aug): 101–115. https://doi.org/10.1016/j.soildyn.2016.05.006.
345–357. https://doi.org/10.1016/j.soildyn.2016.10.009. Tang, C., K.-K. Phoon, and K.-C. Toh. 2014. “Lower-bound limit analysis
Jo, S.-B., J.-G. Ha, M. Yoo, Y. W. Choo, and D.-S. Kim. 2014. “Seismic of seismic passive earth pressure on rigid walls.” Int. J. Geomech. 14 (5):
behavior of an inverted T-shape flexible retaining wall via dynamic cen- 04014022. https://doi.org/10.1061/(ASCE)GM.1943-5622.0000385.
trifuge tests.” Bull. Earthquake Eng. 12 (2): 961–980. https://doi.org/10 Zeng, X. 1998. “Seismic response of gravity quay walls. I: Centrifuge mod-
.1007/s10518-013-9558-9. eling.” J. Geotech. Geoenviron. Eng. 124 (5): 406–417. https://doi.org
Kloukinas, P., A. S. di Santolo, A. Penna, M. Dietz, A. Evangelista, A. L. /10.1061/(ASCE)1090-0241(1998)124:5(406).
Simonelli, C. Taylor, and G. Mylonakis. 2015. “Investigation of seismic Zeng, X., and R. S. Steedman. 2000. “Rotating block method for seismic
response of cantilever retaining walls: Limit analysis vs shaking table displacement of gravity walls.” J. Geotech. Geoenviron. Eng. 126 (8):
testing.” Soil Dyn. Earthquake Eng. 77 (Oct): 432–445. https://doi.org 709–717. https://doi.org/10.1061/(ASCE)1090-0241(2000)126:8(709).
/10.1016/j.soildyn.2015.05.018. Zhou, Y., F. Chen, and X. Wang. 2018. “Seismic active earth pressure for
Madabhushi, S. P. G., and X. Zeng. 1998. “Seismic response of gravity quay inclined rigid retaining walls considering rotation of the principal
walls. II: Numerical modeling.” J. Geotech. Geoenviron. Eng. 124 (5): stresses with pseudo-dynamic method.” Int. J. Geomech. 18 (7):
418–427. https://doi.org/10.1061/(ASCE)1090-0241(1998)124:5(418). 04018083. https://doi.org/10.1061/(ASCE)GM.1943-5622.0001198.

© ASCE 04019117-19 Int. J. Geomech.

Int. J. Geomech., 2019, 19(10): 04019117

You might also like