You are on page 1of 21

Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

www.elsevier.com/locate/soildyn

Effect of structural design on fundamental frequency of reinforced-soil


retaining walls
K. Hatami*, R.J. Bathurst
Civil Engineering Department, Royal Military College of Canada, Kingston, Ont., Canada K7K 7B4
Accepted 5 March 2000

Abstract
The results of a numerical study on the influence of a number of structural design parameters on the fundamental frequency of reinforced-
soil retaining wall models are presented and discussed. The design parameters in the study include the wall height, backfill width,
reinforcement stiffness, reinforcement length, backfill friction angle and toe restraint condition. The intensity of ground motion, character-
ized by peak ground acceleration, is also included in the study as an additional parameter. The study shows that the fundamental frequency of
reinforced-soil wall models with sufficiently wide backfill subjected to moderately strong vibrations can be estimated with reasonable
accuracy from a few available formulae based on linear elastic wave theory using the shear wave speed in the backfill and the wall height.
Numerical analyses showed no significant influence of the reinforcement stiffness, reinforcement length or toe restraint condition on the
fundamental frequency of wall models. The strength of the granular backfill, characterized by its friction angle, also did not show any
observable effect on the fundamental frequency of the reinforced-soil retaining wall. However, the resonance frequencies of wall models
were dependent on the ground motion intensity and to a lesser extent, on the width to height ratio of the backfill. q 2000 Elsevier Science Ltd.
All rights reserved.
Keywords: Fundamental frequency; Reinforced-soil; Retaining walls; Seismic response; Dynamic analysis; Geosynthetics; FLAC

1. Introduction a retaining wall-backfill system is often estimated according to


a one-dimensional shear beam analogy based on the height of
Dynamic lateral earth pressure behind a reinforced-soil the wall and the speed of shear wave in the backfill material
retaining wall subjected to an intensive ground motion can [1,5–10]. In contrast to an infinitely long uniform soil layer, a
be significant. This additional (incremental) horizontal pres- reinforced-soil retaining wall system includes structural
sure may induce excessive wall lateral displacement and components such as reinforcement layers and a vertical-facing
reinforcement load which can result in damage to—or panel supported on a footing. The vertical wall face suggests
collapse of—the structure. Damage to bridge superstruc- that a two-dimensional approach to fundamental frequency
tures, as a result of excessive lateral movement of abutment response analysis may be more appropriate than the one-
retaining walls due to seismic loading has been reported [1–4]. dimensional shear beam approach.
An essential step in seismic design of both conventional and Dynamic response of reinforced-soil retaining walls to
reinforced-soil retaining walls is to determine the natural ground motion has been the subject of several studies [11–
frequencies of the structure. Reinforced-soil retaining walls 16]. However, little can be found in the available literature that
of typical heights (e.g. H , 10 m† and backfill material are specifically addresses the influence of structural design (e.g.
generally considered as short-period structures (e.g. see Ref. reinforcement stiffness, length and spacing, facing panel type
[5]). Soil damping also significantly reduces the contribution and thickness, and toe restraint condition at the panel footing),
of higher modes in total dynamic response of retaining wall geometry and material properties of the backfill, intensity level
systems [6]. Therefore, the response of the wall to ground of shaking and duration of excitation on the fundamental
motion is dominated by the fundamental frequency of the frequency of reinforced-soil retaining wall structures.
structure (also, see Ref. [3]). The fundamental frequency of Richardson and Lee [14] conducted a series of shaking
table studies on small-scale (380 mm high) reinforced-soil
* Corresponding author. Tel.: 1 1-613-541-6000, ext. 6347; fax: 1 1-
wall models. They subjected the retaining wall models to
613-545-8336. harmonic motions with different amplitudes and frequen-
E-mail address: hatami-k@rmc.ca (K. Hatami). cies. The maximum base acceleration varied between 0.02
0267-7261/00/$ - see front matter q 2000 Elsevier Science Ltd. All rights reserved.
PII: S0267-726 1(00)00010-5
138 K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

and 0.50g and the frequency ranged between about 3 and ground acceleration values on the predicted fundamental
40 Hz. The results of frequency sweep of the input base frequency of idealized reinforced-soil wall systems. The
acceleration for each acceleration level provided well- current study reviews closed-form solutions for the predic-
defined frequency response curves. The fundamental tion of the fundamental frequency of one-dimensional and
frequency and the magnitude of acceleration amplification two-dimensional linear elastic media. The results of these
at the surface of the backfill model increased with decreas- closed-form solutions to predict the fundamental frequency
ing input acceleration level. Richardson and Lee concluded of the model retaining walls are compared with values from
that the backfill model responded as a damped, single mode, the results of numerical analyses.
nonlinear elastic oscillator within the examined range of
frequency and acceleration level of the shaking. They also
2. Frequency response analysis of retaining wall models
calculated the fundamental period of various reinforced-soil
wall models using the finite element-based program
2.1. Predicted frequencies from closed-form solutions for
QUAD4B. Based on the results of their finite element simu-
linear elastic soil models
lations they proposed the following empirical equation for
the fundamental period, T1, of a reinforced-soil retaining The fundamental frequencies of the retaining wall models
wall with a level surface: were evaluated based on the backfill soil height and shear
T1 ˆ CH …1† wave speed in order to obtain the appropriate frequency
range for the parametric analysis. The closed-form solutions
where T1 is in seconds, H the height of the wall in meters and provided by Wood [19], Scott [20], Wu [17], and Matsuo
C a coefficient that ranges from 0.020 to 0.033 depending on and Ohara [21] were examined for this purpose.
the shear modulus of the backfill. The two theoretical solutions reported by Matsuo and
Bathurst and Hatami [11,12] carried out numerical simu- Ohara [21] for the fundamental frequency of a linear elastic
lations of the response of 6 m high reinforced-soil wall soil, subjected to horizontal ground motion, are based on
models to variable-amplitude, harmonic input ground two different assumptions: (i) the soil is restricted from any
motions with a range of frequencies. The backfill was vertical displacement …v ˆ 0† throughout the backfill
modeled as a cohesionless, elastic–plastic material with domain and; (ii) no vertical normal stress exists to restrict
Mohr–Coulomb failure criterion. They presented frequency the backfill soil from vertical displacement …s v ˆ 0† over
response plots of the reinforced-soil retaining wall models the entire domain. Matsuo and Ohara argued that the solu-
that were obtained using the calculated maximum wall tion for the real case lies between these two extreme cases.
displacement and reinforcement load. The peak ground They derived the solutions for soil horizontal displacement
acceleration, a g, was set to 0.2g where g is the acceleration and lateral pressure on a rigid wall subjected to harmonic
of gravity. Their results indicated that the fundamental loading for the above two limiting cases in a common
frequencies of reinforced-soil retaining wall models were expression with different coefficients representing each
close to the predicted values based on the conventional case. Their solutions apply to a wall retaining an infinitely
one-dimensional shear beam model for cases with wide wide backfill. The solutions for displacement showed a
backfill. Their results also showed that a proposed formula decay of amplitude with distance from the wall towards
by Wu [17] for a two-dimensional backfill model provided a the far field for loading frequencies below the fundamental
reasonable estimate of the fundamental frequency of the frequency of the soil–wall system. On the other hand, the
reinforced-soil wall models with sufficiently wide backfill, solution indicated radiation of waves towards the far field
B (e.g. B=H . 5†: Hatami and Bathurst [18] also examined for high frequency loading. Matsuo and Ohara found the
the accuracy of the one-dimensional shear beam analogy calculated pressure at the bottom of the wall to be about
and Wu’s two-dimensional solution for retaining wall 10% higher in case (i) than in case (ii) for Poisson’s ratio
fundamental frequency for different wall height and ground n ˆ 0:3:
motion intensity values. They showed that both the one- The fundamental frequencies of the soil–rigid wall
dimensional shear beam equation and Wu’s formula over- system for the two cases described above can be expressed
estimated the fundamental frequencies of the tall walls as follows [17,19,21] (see also, Ref. [22]):
…H ˆ 9 m† under strong ground motion (peak ground accel- Case (i):
eration, ag ˆ 0:4g† whereas both theoretical approaches
resulted in satisfactory frequency predictions for shorter f11vˆ0 ˆ f1 ·GFvˆ0 …2†
…H ˆ 3 m† reinforced-soil wall models. The theoretical
s
predictions were satisfactory for all model heights under 1 G
moderately strong …ag ˆ 0:2g† ground acceleration. f1 ˆ …3†
4H @
The current paper extends the preliminary work of
Hatami and Bathurst [18] by examining the influence of a s
 
wider range of model backfill width to height ratio, soil 8…1 2 n† H 2
GFvˆ0 ˆ 11 …4†
strength (friction angle), reinforcement stiffness and peak 1 2 2n B
K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157 139

Fig. 1. Variation of the geometric factors from different theoretical solutions with normalized backfill width for different values of backfill Poisson’s ratio.

Case (ii): Poisson’s ratio values. The backfill was modeled as a


svˆ0
plane-strain, homogeneous elastic soil. Scott [20] derived
f11 ˆ f 1 ·GFsvˆ0 …5† the equation for natural frequencies of a rigid retaining wall
s assuming the backfill as a one-dimensional shear beam
 
22n H 2 attached to the wall with elastic springs. He calculated the
GF svˆ0 ˆ 11 …6† stiffness of the springs by comparing his equation for funda-
12n B
mental frequency of the retaining wall with the equation
vˆ0 svˆ0
In the above equations, f 11 and f11 are the frequencies given by Wood [19]. This comparison included the effect
(in Hz) of the first (two-dimensional) mode shape of the of backfill width on the calculated frequency of the retaining
elastic medium corresponding to the cases (i) and (ii), wall system. Scott’s equation for the fundamental frequency
respectively; G is the shear modulus; @ the density and n of a two-dimensional soil–wall system with uniform depth
is Poisson’s ratio of the soil. The parameters GFvˆ0 and is given by:
GFsvˆ0 are geometric factors defined here to represent the
two-dimensional effect of a limited-width backfill on the f 11S ˆ f1 ·GFS …7†
fundamental frequency of the soil-wall system. The s
  2
frequency of an infinitely long, uniform soil layer, f1, is 64 1 2 n H
given by Eq. (3). GFS ˆ 1 1 2 …8†
p 1 2 2n B
Wood [19] numerically calculated the roots of the
S
frequency equations associated with the two-dimensional where f11 is the frequency (in Hz) corresponding to the first
boundary value problem of a uniform backfill contained (two-dimensional) mode shape of the backfill medium
between rigid walls. He presented plots of backfill natural (superscript S denotes Scott’s formula) and GFS is the
frequencies as a function of backfill width for different geometric factor according to Scott’s solution.
140 K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

Fig. 2. Variation of the fundamental frequency of retaining walls with normalized backfill width from closed-form and empirical solutions …n ˆ 0:3†:

Wu [17] and Wu and Finn [23,24] developed an approx- (two-dimensional) mode shape of the soil medium (super-
imate closed-form solution for dynamic earth pressure on a script W denotes Wu’s formula) and GFW is the geometric
rigid wall with the assumption of a homogenous, elastic factor according to Wu’s solution.
backfill. Their formulation included approximations based The variation of geometric factors with normalized width
on shear beam analogy with zero shear stress at the surface of the backfill …B=H† from theoretical solutions (i.e. Eqs. (4),
and assumption of no vertical normal stress throughout the (6), (8), and (10) and the reproduced plots of Wood [19]) is
backfill. The wall-backfill system was assumed to be a plotted in Fig. 1.
plane-strain model. The wall and the foundation were both It is seen that for the case of an infinitely wide backfill
assumed to be rigid. Accordingly, the displacement field …B=H ! ∞†; the geometric factors based on all the above
across the backfill could be approximated as a summation solutions approach unity and the corresponding frequencies
of sinusoidal mode shapes in horizontal and vertical direc- converge to f1. It is also seen that the predicted fundamental
tions. Wu [17] derived a closed-form expression for the frequency of the wall for different values of B=H and Pois-
undamped natural frequencies of the backfill model subject son’s ratio according to all the above solutions fall between
to the above boundary conditions and under small amplitude the predicted values based on the two limiting cases of v ˆ
vibration. Wu argued that the nonlinear behavior of the 0 and s v ˆ 0: The difference between the two cases v ˆ 0
backfill under strong ground motion would reduce the and s v ˆ 0 increases significantly as n approaches the value
fundamental frequency of the soil-wall system and therefore 0.5. According to Fig. 1, the v ˆ 0 approximation results in
would alter the peak seismic thrust on the wall compared to a relatively large predicted frequency of the soil-wall
the case of a linear soil model. However, no quantitative system for n close to 0.5. However, for Poisson’s ratio
evaluation of the expected difference in the response was values of typical granular soils used in reinforced-soil
provided. Wu suggested that his analysis could be extended retaining walls …n # 0:3† and sufficiently wide backfill
to stronger input ground motions by evaluating the reduced (e.g. B=H . 5†; the above limiting approximations show
shear modulus of the backfill at larger strain levels and almost no difference in predicted values for the fundamental
calculating the modified fundamental frequency of the frequency of soil-retaining wall systems. The value of GF
soil-retaining wall system. The fundamental frequency of using Scott’s solution also increases significantly for large
the soil model behind a rigid wall under small-amplitude values of Poisson’s ratio …n ! 0:5† and narrow backfill (e.g.
vibrations according to Wu [17] is given by: B=H # 3† (Fig. 1). Large Poisson’s ratio values and narrow
backfill result in strong two-dimensional effects which
W
f11 ˆ f1 ·GFW …9† violate the v ˆ 0 condition and the one-dimensional approx-
s imation used by Scott (shear beam analogy). Fig. 1 also
  2 shows that Scott’s prediction of the retaining wall funda-
2 H
GFW ˆ 1 1 …10† mental frequency is very close to the prediction based on the
12n B
v ˆ 0 approximation. The close agreement can be expected
W
where f11 is the frequency (in Hz) corresponding to the first since Scott [20] used the same approximation to determine
K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157 141

Fig. 3. Example numerical grid for reinforced-soil wall with fixed toe condition.

the stiffness of the springs representing the soil–wall inter- Richardson and Lee [14]. However, Eq. (11) is valid for a
action. particular sandy soil compacted to medium relative density
Non-dimensionalized fundamental frequencies, f11 =Hvs ; and may not apply for other soils. In addition, the formula
for retaining wall models with different backfill width are does not explicitly include the effect of backfill aspect ratio
plotted in Fig. 2 and have been calculated using the (width to height) on the predicted fundamental frequency,
geometric factors shown in Fig. 1 (case of n ˆ 0:3†: The which may be significant for a narrower backfill as shown in
backfill properties assumed for numerical models of this Figs. 1 and 2. Fairless [26] reviewed the results of physical
study (described later in section 2.4.) are used to calculate tests reported in the literature and demonstrated that Eq.
the shear wave speed, vs. The two limiting cases v ˆ 0 and (11) typically overestimated the measured fundamental
s v ˆ 0 are not included in this figure for brevity. However, frequencies of model reinforced-soil retaining walls.
they would show the same relative trends as in Fig. 1 with Overall, Richardson’s empirical estimation of the funda-
respect to other solutions. Fig. 2 also includes the non- mental frequency of reinforced-soil retaining walls is within
dimensionalized form of the fundamental frequency of the range of predicted results using the theoretical solutions
retaining wall models using the empirical formula proposed for conventional retaining walls (Fig. 2) but may result
by Richardson [25]: in values significantly greater than the fundamental
frequency values predicted using the theoretical solutions
38:1
f11R ˆ …11† with B=H . 3:
H The above theoretical predictions for the fundamental
R
where f11 is the estimated fundamental frequency of the frequency of soil-retaining wall systems are based on linear
reinforced-soil wall in Hz (superscript R denotes Richard- theory and do not include the influence of the shaking inten-
son’s formula) and H is the wall height in meters. The sity level which can generate nonlinear and plastic response
empirical formula in Eq. (11) was proposed based on the of the structure. Other attempts to predict the fundamental
measured response of a full-scale (6.1 m high) test wall frequency of gravity and earth dam structures are also cited
excited by a buried explosive charge and forced vibration in the literature that are valid for low-amplitude ground
tests of four existing reinforced-soil walls of different motions where linear elastic theory is applicable [27].
heights ranging from 2.3 to 8.5 m [8,25]. In following text, the results of theoretical solutions
Richardson et al. [8] showed a good agreement between described in this section are compared against the results
Eqs. (3) and (11) using their estimate of the backfill shear of numerical evaluation of the fundamental frequency of
modulus values of their wall models. Hence, Eq. (11) reinforced-soil retaining wall models subjected to different
provided a simple and satisfactory estimation of the funda- ground motion intensities.
mental frequency of the reinforced-soil walls that were
tested. The predicted fundamental frequency of rein- 2.2. Numerical approach
forced-soil retaining walls based on Eq. (11) is consistent
with the suggested range in Eq. (1) proposed earlier by The two-dimensional, finite difference program Fast
142 K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

Table 1
Parametric values used in the evaluation of the fundamental frequency of wall models (notes: reinforcement spacing Sv ˆ 1:0 m; viscous damping ratio j ˆ
5%†

Wall height Model width Width to height Reinforcement Toe Restraint Reinforcement Friction Peak ground Input frequency
(m) B (m) ratio B=H ratio L=H condition stiffness J (kN/m) angle (8) acceleration a g (g) fg (Hz)

3 30 10 0.4, 1.0 Fixed 500, 10 000, 69 000 0.2, 0.4 3, 5, 6, 7


6 42 7 0.4, 1.0 Fixed, Sliding 500, 10 000, 69 000 20, 35, 45 0.2, 0.4 1, 2.5, 3, 3.5, 4
9 18, 36, 54 a 2, 4, 6 0.4, 1.0 Fixed 500, 10 000, 69 000 0.05, 0.2, 0.4 1, 1.5, 2, 2.5, 3, 4
a
Reference wide-width backfill model case.

Fig. 4. Variation of normalized maximum lateral displacement of wall crest with normalized frequency of input ground motion: (a) J ˆ 500 kN/m;
(b) J ˆ 10000 kN/m; (c) J ˆ 69000 kN/m.
K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157 143

Fig. 4. (continued)

Lagrangian Analysis of Continua (FLAC 3.40—[28]) was the back of the facing panel (Table 1) so as to contain the
used to carry out the numerical experiments. The program is shear wedge (plastic zone) that develops behind the rein-
widely used in geotechnical engineering applications and is forced zone during base shaking. The vertical spacing
attractive for seismic analysis of reinforced-soil retaining between reinforcement layers was kept constant at Sv ˆ
walls because it can model large distortions and near- 1:0 m: The height of the wall models …H ˆ 3; 6 and 9 m)
collapse conditions [11,12,29]. and the spacing between reinforcement layers are typical of
actual structures in the field. The toe restraint condition
2.3. Numerical grid and problem boundaries (wall footing) was either fixed (i.e. the toe of the wall was
slaved to the foundation but was free to rotate) or free to
The numerical grid for a typical wall model used in the slide horizontally and rotate about the toe. The results of a
study is illustrated in Fig. 3. Numerical simulations repre- previous study [11] show that the lateral displacement of the
sent a backfill of constant depth retained by a continuous wall and the magnitude and distribution of reinforcement
panel wall with uniformly spaced reinforcement layers. The load can be significantly affected by the wall toe restraint
width of the backfill, B, in the cases representing an infi- condition. For sliding cases, the wall model was seated on a
nitely wide backfill was extended to a large distance beyond thin (0.05 m thick) layer of soil that was extended across the
144 K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

Fig. 4. (continued)

entire width of the numerical grid. This layer performed a L=H ˆ 0:33 and 0.4 have been used in some shaking table
similar function to a sliding interface and was required to studies [30–33].
ensure that models representing walls without horizontal toe
restraint (i.e. sliding-wall cases) were not artificially 2.4. Material properties
restrained along their contact area with the foundation
during shaking. For the fixed-toe condition, the wall and The wall facing was modeled as a continuous concrete
soil regions were connected directly to a foundation base panel with a thickness of 0.14 m. The bulk and shear modu-
comprising of a 1m-thick layer of very stiff material (Fig. 3). lus values of the wall were Kw ˆ 11 430 MPa and Gw ˆ
Two values for the reinforcement length to wall height ratio 10 430 MPa†; respectively. Poisson’s ratio for the panel
were selected to represent narrow …L=H ˆ 0:4† and wide material was taken as nw ˆ 0:15: The granular backfill
…L=H ˆ 1† reinforced zones in the parametric analysis. was modeled as a purely frictional, elastic–plastic soil
These reinforcement ratio values capture the range of values with a Mohr–Coulomb failure criterion. The reference fric-
reported in the literature for actual structures in the field as tion angle of the soil was f ˆ 358; dilatancy angle c ˆ 68;
well as experimental studies on reduced-scale retaining wall and unit weight g ˆ 20 kN=m 3 : The soil material was
models. Reinforcement length to wall height ratios as low as assigned constant values of bulk modulus K s ˆ 27:5 MPa
K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157 145

Fig. 5. Variation of response ratio of maximum wall lateral displacement (response ratio between ag ˆ 0:2g and ag ˆ 0:4g loadings) with normalized
frequency of input ground motion.

and shear modulus Gs ˆ 12:7 MPa: The foundation zone for 69 000 kN/m (Table 1). The lower, intermediate and higher
fixed-toe cases was assigned the same material properties as stiffness values represent an extensible (polymeric) geotex-
the concrete facing panel. The panel-soil interface was tile reinforcement, a very stiff (polymeric) geogrid reinfor-
modeled using a thin (0.05 m thick) soil column directly cement material and a steel strip reinforcement,
behind the facing panel. The friction angle and the dilatancy respectively. A large range of stiffness values was included
angle of the interface soil column between the reinforced in the parametric study to identify any possible stiffness
zone and the facing panel were set to fi ˆ 208 and c i ˆ 0; effects on the resulting resonance frequency of the retaining
respectively. The remaining soil properties of the interface walls with the input ground motion. The yield strength of
soil column were the same as the properties of the backfill the reinforcement in all cases was kept constant at Ty ˆ
soil. The reinforcement layers were modeled using linear, 200 kN=m; which is well above the magnitude of the maxi-
elastic–plastic cable elements with negligible compressive mum reinforcement load recorded in the simulations.
strength and an equivalent cross-sectional area of 0.002 m 2. Consequently, reinforcement rupture was not a possible fail-
The equivalent linear elastic stiffness values for the reinfor- ure mechanism in this study. The interface between the
cement layers were taken as J ˆ 500; 10 000 and reinforcement (cable elements) and the soil was modeled
146 K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

Fig. 6. Variation of maximum reinforcement incremental load with normalized frequency of input ground motion: (a) J ˆ 500 kN/m; (b) J ˆ 10000 kN/m;
(c) J ˆ 69000 kN/m.

with a grout material of negligible thickness and with an while the continuous wall facing panel was braced horizon-
interface friction angle dg ˆ 358: The grout’s bond stiffness tally using rigid external supports. The panel supports were
and bond strength values were taken as kb ˆ 2 × 10 3 MN/ then released in sequence from the top to the bottom of the
m/m and sb ˆ 1 × 103 kN=m; respectively. These interface structure as is done in the field. After static equilibrium was
and grout properties were selected to simulate a perfect achieved (end of construction stage), the full width of the
bond between the soil and reinforcement layers. The end foundation was subjected to the variable-amplitude harmo-
of each cable element was connected to a single grid point at nic ground motion record illustrated in Fig. 3 (inset). This
the back surface of the facing panel region to simulate a acceleration record was applied horizontally to all nodes at
fixed reinforcement connection in the field. the bottom and the right-hand side (truncated) boundary of
the backfill region at equal time intervals of Dt ˆ 0:05 s:
The mathematical expression for input acceleration is
2.5. Seismic loading
given by:
The staged construction of each wall model was simu- q
lated by placing the backfill and the reinforcement in layers  ˆ b e2at tz sin…2pft†
u…t† …12†
K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157 147

Fig. 6. (continued)

where a ˆ 5:5; b ˆ 55; and z ˆ 12 are constant the vertical truncated boundary was based on the assump-
coefficients, f is the base acceleration frequency and, t is tion of uniform distribution of horizontal acceleration over
the time. The resulting peak amplitude, a g, using these the depth of the backfill away from the facing panel. This
parameters is 0.2g, where g is the acceleration of gravity. assumption also represents the case of reduced-scale rein-
Coefficient terms were adjusted to give a peak amplitude forced-soil retaining wall models on shaking tables where a
ag ˆ 0:4g representing a stronger earthquake. The variable- rigid back boundary moves in phase with the base [14]. A
amplitude input ground motion from Eq. (12) was chosen viscous damping ratio of j ˆ 5% was chosen for both the
over simple harmonic acceleration because it simulates the soil and facing panel regions in the parametric analyses.
rise and time decay of an idealized accelerogram. In addi- This damping ratio value may appear conservative but it
tion, it does not lead to excessive response of the wall model is comparable to the range of values estimated based on
that can be expected in the case of a sustained, constant measured response of retaining walls to dynamic loading
amplitude input base acceleration. The input acceleration in a number of experimental studies [5,8]. In addition, a
in Eq. (12) is similar to the Tsang signal function reported major portion of input seismic energy is dissipated
elsewhere [34]. The application of uniform acceleration at through the hysteresis and plastic deformation of the soil.
148 K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

Fig. 6. (continued)

A relatively low value for the backfill damping ratio …j ˆ 3.1. Effect of reinforcement length and stiffness
5%† also ensures a detectable difference in dynamic
response of retaining wall models for different parametric 3.1.1. Wall displacements
cases. Fig. 4 summarizes the variation of normalized maximum
lateral displacement of the wall crest with normalized
frequency of input ground motion. The datum for all displa-
3. Response of wall models to input ground motion cement plots in this study is taken with respect to the end-of-
construction condition following prop release. The toe
The calculated response of wall models to the introduced restraint condition is fixed in all the cases presented in
base motion (Eq. (12)) is represented in terms of maximum Fig. 4. The input frequency is nondimensionalized in the
normalized lateral displacement of the facing panel and form v H/vs where v is the input circular frequency and vs
maximum reinforcement load during shaking. These para- is the speed of shear wave propagation in the backfill
meters are selected as representative parameters in the material.
present frequency response analysis due to their significance Progressive outward displacement of the wall facing
in design and performance of retaining wall structures under was observed during all the parametric simulation runs.
both static and seismic loading conditions. The accumulated, dynamic portion of the wall lateral
K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157 149

Fig. 7. Variation of response ratio of maximum reinforcement incremental load (response ratio between ag ˆ 0:2g and ag ˆ 0:4g loadings) with normalized
frequency of input ground motion.

displacement was significant in all analysis cases. A shear estimate the numerical results. This overestimation is more
failure wedge developed and extended well beyond the rein- significant at lower B=H ratio values (e.g. compare theore-
forced zone. The observed wedge angle, in general, showed tical predictions in the plots for cases H ˆ 3 m (where
good agreement with the predicted value according to the B=H ˆ 10† and H ˆ 9 m (where B=H ˆ 6† in Fig. 4). The
Mononabe–Okabe theory considering amplification of predicted fundamental frequencies of wall models using any
acceleration over the height of the wall [11]. Despite signif- of the linear elastic theories become less accurate for the
icant wall displacement and evidence of a plastic region in stronger ground motion ag ˆ 0:4g: The normalized
the backfill, the frequency response curves in Fig. 4 indicate frequency value according to Richardson’s empirical equa-
that the fundamental frequency of wall models with wide tion (Eq. (11)) is v11 H=vs ˆ 3:03 which clearly overesti-
backfill (e.g. B=H . 5† subjected to moderately strong mates the fundamental frequency of the retaining wall
ground motion …a g ˆ 0:2g† is predicted satisfactorily models plotted in Fig. 4.
using one-dimensional theory (Eq. (3)) and Wu’s solution Comparison of the frequency response curves shown in
(Eqs. (9) and (10)). The fundamental frequencies of wall Fig. 4a–c reveals almost no influence of the reinforcement
models based on Wood and Scott solutions generally over- stiffness or length on the predicted low-amplitude
150 K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

Fig. 8. Influence of toe restraint condition on frequency response of wall to input ground motion …H ˆ 6 m; L=H ˆ 1; ag ˆ 0:4g†:

fundamental frequency of a wall model of given height and due to the short width of the reinforced zone. A docu-
backfill material. This conclusion is consistent with obser- mented case of translational slip at the base of a rein-
vations by Wolfe [16] who found that for a given height of forced-soil wall in Japan during the Kobe earthquake
model wall tested on a shaking table, the reinforcement has been attributed to insufficient reinforcement length
density did not influence the observed low-amplitude funda- [36].
mental frequency of the structure (also, see Ref. [26]). In the current parametric analyses, some cases of instabil-
However, the actual magnitude of wall displacement is ity in numerical models with L=H ˆ 0:4 and J ˆ 500 kN=m
sensitive to reinforcement stiffness value and reinforcement subjected to ag ˆ 0:4g ground motion were observed (Fig.
length as demonstrated in parametric analyses presented by 4a). These cases are also consistent with observations by
Bathurst and Hatami [11]. Wolfe [16] in shaking table studies of model reinforced-
Compared to the cases with L=H ˆ 1:0; greater displace- soil walls where large lateral displacements of models
ments were observed in wall models with short reinforce- with short reinforcement length under strong shaking were
ment length …L=H ˆ 0:4† under strong …ag ˆ 0:4g†; low- observed. Wolfe noted that the wall response is significantly
frequency input ground motion (Fig. 4). This is mainly more sensitive to the loading intensity than the loading
due to insufficient reinforcement length for the case L=H ˆ frequency when the predominant frequency of the input
0:4: Design codes such as FHWA [35] recommend a mini- motion is lower than the fundamental frequency of the
mum L=H ˆ 0:7 for the reinforcement length in reinforced- retaining wall. Wolfe suggested that a minimum reinforce-
soil walls under static loading. The wall response to input ment length can be determined by treating the reinforced
ground motion with a frequency lower than the wall soil zone as a gravity mass and determining the width of the
fundamental frequency is essentially a quasi-static gravity block required to resist dynamic lateral earth forces
response to lateral earth pressure. Accordingly, a wall acting against the sliding mass and inertial force of the mass.
with under-designed reinforcement length may undergo This approach has been adopted in recent Newmark-type
significant lateral displacement and possible instability [37] and seismic pseudo-static design methods for
K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157 151

Fig. 9. Variation of maximum wall response with normalized frequency of input ground motion …H ˆ 6 m; L=H ˆ 1; ag ˆ 0:2g; fixed-toe condition):
(a) normalized lateral displacement of wall crest; (b) reinforcement load.

reinforced–soil walls constructed with modular block the response of a purely linear elastic system, in which the
facings [38–40]. response magnitude is proportional to the input intensity
Fig. 5 shows another representation of the variation of the level, the plastic deformation of the backfill at resonance
wall displacement response with input frequency. The ratio is great enough to control the system response for both
of the calculated wall lateral displacement under the two values of input ground motion intensity used in the current
ground motion intensity levels ag ˆ 0:2g and 0.4g generally study.
depends on both the ratio of input intensity levels and
the input frequency. The influence of the input loading 3.1.2. Reinforcement loads
intensity approaches a minimum at the resonance frequency Fig. 6 summarizes the variation of maximum dynamic
of the retaining wall with ground motion. This phenomenon reinforcement load, Tmax, with normalized loading
is clearly observed in Fig. 5 which is consistent with the frequency. The reported reinforcement forces are incremen-
fundamental frequency of the wall models inferred from tal values taken with respect to values calculated at the end
Fig. 4. The minimum response ratio value of about one in of construction and following prop release. The maximum
Fig. 5 is also another indication of nonlinearity and plastic load in each reinforcement layer was observed at the
response of the retaining wall system when subjected to connection point to the facing panel. However, contrary to
moderately strong to strong ground motion. In contrast to values for wall lateral displacement that was the largest at
152 K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

Fig. 9. (continued)

the wall crest, the layer with the maximum reinforcement frequency of the retaining wall with ground motion in a
incremental load varied between parametric cases. This may similar fashion to that observed for the data in Fig. 5.
explain why the frequency response curves of wall lateral
displacement (Fig. 4) are better than the corresponding rein- 3.2. Influence of toe restraint condition
forcement load plots (Fig. 6) with respect to defining the
Frequency responses of relative lateral displacement of
fundamental frequency value of the retaining wall struc-
the wall crest and maximum reinforcement incremental load
tures. Typically the maximum incremental load, Tmax, was
for the two cases of fixed-toe and sliding-toe condition are
observed in lower reinforcement layers for stiffer reinforce-
shown in Fig. 8. The results are shown for 6m-high wall
ment, shorter reinforcement length and, stronger input
models with L=H ˆ 1 and ag ˆ 0:4g: The stronger input
ground motion for a given set of other parameters. Cases
ground motion …ag ˆ 0:4g† was chosen in order to magnify
with L=H ˆ 0:4 resulted in large incremental load at lower
any possible influence of the toe restraint condition on the
frequencies which can be attributed to the under-designed
fundamental frequency of the wall. The frequency responses
reinforcement length and quasi-static loading condition.
of lateral displacement and reinforcement load do not show
Nonetheless, the frequency response of the incremental
any observable dependence of fundamental frequency on
load also shows a maximum (with better accuracy for
toe restraint condition of the wall models.
moderately strong ground motion of ag ˆ 0:2g† in the vici-
nity of the predicted fundamental frequency based on linear- 3.3. Influence of soil friction angle
elastic analysis. The observations noted earlier regarding the
accuracy of theoretical solutions to predict wall lateral Fig. 9a shows the calculated frequency response of wall
displacement frequency response (Fig. 4) are also applicable lateral displacement for three different values of friction
to frequency response of reinforcement load. angle, f , for the backfill soil. Two reinforcement stiffness
Fig. 7 shows the variation of reinforcement load response cases were investigated: J ˆ 500 kN=m representing a poly-
ratio with input frequency. The influence of the input load- meric reinforcement material and, J ˆ 69 000 kN=m repre-
ing intensity approaches a minimum at the resonance senting a metallic reinforcement product. The calculated
K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157 153

Fig. 10. Variation of maximum wall response with normalized frequency of input ground motion for different backfill width cases
…H ˆ 9 m; L=H ˆ 1; J ˆ 10 000 kN=m†: (a) normalized lateral displacement of wall crest; (b) reinforcement load.

maximum dynamic lateral displacement of the wall gener- shown for 9m-high wall models with L=H ˆ 1:0 and J ˆ
ally increases with decreasing value of backfill friction 10 000 kN=m:
angle and lower reinforcement stiffness. However, neither The predictions of the fundamental frequency of wall
of the above parameters shows a measurable influence on models in Fig. 10 using two–dimensional solutions show
the predicted fundamental frequency of wall displacement significant differences from the prediction based on one-
for the reinforced-soil wall models. A similar conclusion dimensional elastic theory for the case of a narrow backfill
can be made with respect to frequency response of maxi- (e.g. B=H ˆ 2†: The prediction of fundamental frequency
mum reinforcement load data in Fig. 9b. Consistent with the based on Wu’s approach shows the closest agreement with
results of Fig. 4, excessive lateral displacement in the lower the frequency values inferred from the numerical results in
frequency range was observed for cases with low friction Fig. 10 of all the two-dimensional solutions evaluated. The
angle as a result of excessive failure of the backfill soil agreement is most satisfactory for small-amplitude vibra-
(plasticity). However, all analyses were numerically stable. tions (i.e. case of ag ˆ 0:05g† and is better than the one-
dimensional prediction for the retaining walls with a narrow
3.4. Influence of backfill width and ground motion intensity backfill (Fig. 10a and b with B=H ˆ 2†: The theoretical
predictions of the fundamental frequency of wall models
Frequency responses of normalized lateral displacement (vertical lines in Fig. 10a and b) for the cases B=H ˆ 2
of the wall crest and maximum reinforcement incremental and 4 show a strong dependence on the backfill B=H ratio
load for the two backfill aspect ratio values B=H ˆ 2 and 4 (consistent with the plots in Figs. 2 and 4). The fundamental
are given in Fig. 10a and b, respectively. The results are frequency values inferred from large response amplitude
154 K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

Fig. 10. (continued)

regions in the plots of numerical results show much less nonlinear characteristic of the structure under severe excita-
influence of the B/H ratio than the closed-form solution tion. This shift of frequency was more pronounced for taller
results, particularly for the stronger input ground motions. wall models. This may be attributed to larger strain levels
The fundamental frequencies of the wall models from that are developed in 9 m high models as compared to the
numerical results consistently shift toward lower frequen- 3 m high models. The frequency response curves of wall
cies under stronger input ground motions (i.e. ag ˆ 0:2g lateral displacement and reinforcement load show well-
and 0:4g compared to ag ˆ 0:05g; also, compare cases defined peak characteristics for all wall heights and reinfor-
of 0.4g and 0.2g in Fig. 4) and happen to approach the cement stiffness values, which is in accordance with the
predicted value based on the one-dimensional solution results of shaking table studies reported by Richardson
(Eq. (3)). This observation is attributed to the decrease in and Lee [14].
the magnitude of the modulus of the backfill material at The response of the reinforced-soil retaining wall models
larger strain levels which has also been observed in experi- in the vicinity of resonance (i.e. proximity to v ˆ p vs =2H†
mental studies [14]. The equivalent viscous damping ratio was difficult to calculate due to numerical stability
of the backfill increases with increasing strain level. A larger problems. In addition, due to plastic behavior and excessive
damping ratio value also reduces the fundamental frequency local deformation of the backfill at the surface away from
of the retaining wall structure under strong ground motion to the wall crest, the calculated lateral displacement at the wall
values less than the predicted value based on undamped, crest does not increase in the vicinity of the wall resonance
linear elastic analysis. The dependence of fundamental frequency with increasing ground motion intensity. Accord-
frequency of the wall on ground motion intensity is a ingly, the displacement response values very close to the
K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157 155

resonance frequency cannot be reliably used to complete the based on one-dimensional theory for significantly wide
well-defined frequency response curves by numerical calcu- backfill (e.g. B=H . 10†:
lations. However, some theoretical studies on conventional Earlier numerical simulation work by the writers [11] has
retaining walls [6,41] indicate a resonance amplification demonstrated that reinforcement stiffness, reinforcement
factor of (2j ) 21/2 in wall response compared to the value length and toe restraint condition can have a significant
(2j ) 21 which corresponds to a simple linear oscillator with influence on the magnitude of reinforcement forces and
viscous damping ratio, j . For j ˆ 5%; the magnitude of lateral displacements of reinforced-soil wall models during
amplification factor at resonance will be about three a simulated seismic event. However, the results of the
which is considerably less than the value of 10 associated current study using the same numerical models demonstrate
with a viscously damped linear oscillator. that these variables do not significantly affect the fundamen-
tal frequency of reinforced-soil wall models with a wide
range of structural component values.
4. Conclusions Large response magnitude was observed at low frequen-
cies for the cases where the reinforcement length, reinforce-
Retaining walls of typical heights (e.g. H , 10 m† are ment stiffness or backfill friction angle was very low. The
considered as short-period structures and therefore, their seis- fundamental frequency of a retaining wall with narrow
mic response is dominated by their fundamental frequency. backfill according to theoretical predictions can be signifi-
The paper first summarizes theoretical solutions for evaluating cantly higher than the case of a wall with an infinitely wide
the fundamental frequency of conventional retaining walls backfill. However, the numerical results of this study on
(typically rigid retaining walls). These solutions can be fundamental frequency of model reinforced-soil walls
applied to continuum, plane-strain models of retaining wall- were relatively less sensitive to the backfill width of the
backfill systems and are presented in the general form: reinforced-soil wall system. The reason for the reduced
f 11 ˆ f1 ·GF …13† effect of the backfill width is attributed to soil plasticity
that develops in the near-field behind the facing panel. In
where GF ˆ f …n; B=H† represents the modification of f1 to contrast, purely elastic soil response is more significantly
obtain the fundamental frequency of a two-dimensional retain- influenced by the geometry, i.e. backfill width, of the model.
ing wall model from the one-dimensional frequency formula Another reason for the difference may be partly due to the
for an infinitely long uniform soil layer. type of boundary condition at the truncated far-end bound-
The results of theoretical solutions are compared to ary of the backfill. The uniform acceleration applied coher-
the results of numerical modeling of a wide range of ently with the base acceleration at the truncated far-end
reinforced-soil retaining wall models subjected to base boundary of reinforced-soil wall models in this study may
excitation using a variable-amplitude harmonic input subdue the influence of backfill width compared to the case
acceleration record with a range of frequencies in the with a nonaccelerated, radiating boundary. However, since
vicinity of the predicted values according to linear elastic the same far-end boundary condition was used in all
analysis. Parametric seismic analyses on reinforced-soil parametric cases, the type of boundary condition does
retaining wall models were carried out to investigate the influ- not effect the relative response of the different models
ence of different structural components on their fundamental investigated.
frequency. The structural components included the reinforce- It is also noted that the available theoretical approaches
ment stiffness and length, the restraining condition at the toe do not result in satisfactory estimation of the fundamental
(footing) of the facing panel and the friction angle of frequency of retaining walls with limited-width backfill for
the granular backfill soil. Problem geometry parameters the case where the retaining wall structure is expected to
included the wall height and the backfill width. The inten- experience a severe ground motion. The intensity of the
sity of ground motion, characterized by the peak ground input ground motion, represented by the magnitude of the
acceleration, was also varied. peak acceleration showed the most dominant influence on
The results of the analyses showed that the fundamental the resonance frequency of a given retaining wall model.
frequency of reinforced-soil retaining wall systems with a The resonance frequencies were lower under a stronger
sufficiently wide uniform backfill subjected to moderately input acceleration and were different from the predicted
strong ground motion (e.g. ag ˆ 0:2g in the present study) values based on low-amplitude, linear elastic analysis.
can be estimated with reasonable accuracy from a Also, the difference between numerical and theoretical
commonly used one-dimensional solution based on linear predictions of fundamental frequency is larger for retaining
elastic theory. Among the two-dimensional approaches wall models subjected to stronger input ground motion.
examined, the frequency formula proposed by Wu [17]
and Wu and Finn [23,24] gave the closest agreement to
the fundamental frequency value inferred from numerical 5. Additional remarks
results. The fundamental frequency values from two-dimen-
sional continuum models were shown to approach values The conclusions of this study are limited to model
156 K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157

retaining walls with uniform backfill soils and constant Recent Advances in Geotechnical Earthquake Engineering and Soil
material properties. Nonlinear effects such as interface slip Dynamics, St. Louis, Missouri, 1995. p. 1589–604.
[7] Al-Homoud AS, Whitman RV. Evaluating tilt of gravity retaining
or rupture of the reinforcement and stress-dependent proper- walls during earthquakes. Proceedings of the 10th World Conference
ties of soil were not addressed in this study. Stress-depen- on Earthquake Engineering. Madrid, Spain, vol. III, 1992. p. 1683–8.
dent (e.g. hyperbolic model proposed by Duncan et al. [42]) [8] Richardson GN, Feger D, Fong A, Lee KL. Seismic testing of rein-
or strain-dependent modulus values of the backfill as forced-earth walls. Proceedings of ASCE, Journal of the Geotechnical
proposed by Seed and Idriss [8] may result in an even Engineering Division 1977;103(GT1):1–17.
[9] Siddarthan R, Norris GM. On the seismic displacement response of
greater influence of ground motion intensity on the reso- rigid retaining walls. Soils and Foundations 1991;31(2):51–64.
nance frequency of retaining wall systems than reported [10] Whitman RV. Seismic design and behavior of gravity retaining walls.
here. Richardson [25] introduced a frequency correction Proceedings of Specialty Conference on Design and Performance of
factor, FCF, to modify the predicted fundamental frequency Earth-Retaining Structures. ASCE Special Publication No. 25, 1990.
of reinforced-soil walls from linear elastic theory according p. 817–42.
[11] Bathurst RJ, Hatami K. Seismic response analysis of a geosynthetic-
to the expected peak dynamic strain amplitude in the retain-
reinforced soil retaining wall. Geosynthetics International 1998;5(1-
ing wall-backfill system. A strain-dependent stiffness model 2):127–66.
for the backfill and further development of frequency [12] Bathurst RJ, Hatami K. Influence of reinforcement properties on seis-
correction factors are reserved for future study. mic response and design of reinforced-soil retaining walls. Proceed-
Finally, the intensity of the input ground motion is char- ings of the 51st Canadian Geotechnical Conference, Edmonton,
Alberta, vol. II, 1998. p. 479–86.
acterized in this study by its peak acceleration value. Alter-
[13] Cai Z, Bathurst RJ. Seismic response analysis of geosynthetic rein-
natively, ground motion intensity can be represented by forced soil segmental retaining walls by finite element method.
other intensity parameters (e.g. as defined by Arias [43] or Computers and Geotechnics 1995;17:523–46.
spectral intensity defined by Housner [44]). These effects [14] Richardson GN, Lee KL. Seismic design of reinforced earth walls.
are believed to have a quantitative influence on the response Journal of the Geotechnical Engineering Division
1975;101(GT2):167–88.
of retaining wall systems to ground motion. However, the
[15] Segrestin P, Bastick MJ. Seismic design of reinforced-earth (retaining
major conclusion of this study regarding the negligible walls—the contribution of finite element analysis. International
influence of structural component values on the predicted Geotechnical Symposium on Theory and Practice of Earth Reinforce-
resonance frequency of reinforced-soil retaining wall ment, Fukuoka, Japan, 1988. p. 103–47.
systems is not likely to be changed. [16] Wolfe WE. Earthquake-induced deformations in reinforced-earth
walls. PhD thesis, UCLA, Los Angeles, California, 1979.
[17] Wu G. Dynamic soil-structure interaction: pile foundations and
retaining structures. PhD thesis, University of British Columbia,
Acknowledgements Vancouver, Canada, 1994.
[18] Hatami K, Bathurst RJ. Frequency response analysis of reinforced-
soil retaining walls. Proceedings of the 8th Canadian conference on
The funding for the work reported in the paper was
Earthquake Engineering, Vancouver, 1999. p. 341–6.
provided by grants from the Natural Sciences and Engineer- [19] Wood JH. Earthquake-induced earth pressures on structures. Report
ing Research Council and Department of National Defence No. EERL 73-05, California Institute of Technology, Pasadena, Cali-
(Canada). fornia, 1973.
[20] Scott RF. Earthquake-induced earth pressures on retaining walls.
Proceedings of the 5th World Conference on Earthquake Engineering,
Rome, Italy, vol. II, 1973. p. 1611–20.
References [21] Matsuo H, Ohara S. Lateral earth pressure and stability of quay walls
during earthquakes. Proceedings of the 2nd World Conference on
[1] Al-Homoud AS, Whitman RV. Seismic analysis and design of Earthquake Engineering, Tokyo-Kyoto, Japan, vol. I, 1960. p. 165–
rigid bridge abutments considering rotation and sliding incorporating 81.
non-linear soil behavior. Soil Dynamics and Earthquake Engineering [22] Ishii Y, Arai H, Tsuchida H. Lateral earth pressure in an earthquake.
1999;18:247–77. Proceedings of the 2nd World Conference on Earthquake Engineer-
[2] Bakeer RM, Bhatia SK, Ishibashi I. Dynamic earth pressure with ing, Tokyo-Kyoto, Japan, vol. I, 1960. p. 211–30.
various gravity wall movements. Proceedings of ASCE Specialty [23] Wu G, Liam Finn WD. Seismic pressures against rigid walls.
Conference on Design and Performance of Earth-Retaining Struc- Proceedings of the ASCE Specialty Conference on Analysis and
tures, Ithaca, NY, 1990. p. 887–99. Design of Retaining Structures against Earthquakes. Geotechnical
[3] Oritz LA, Scott RF, Lee J. Dynamic centrifuge testing of a cantilever Special Publication No. 60, 1996. p. 1–18.
retaining wall. Earthquake Engineering and Structural Dynamics [24] Wu G, Liam Finn WD. Seismic lateral pressures for design of rigid
1983;11:251–68. walls. Canadian Geotechnical Journal 1999;36:509–22.
[4] Seed HB, Whitman RV. Design of retaining structures for dynamic [25] Richardson GN. Earthquake resistant reinforced-earth walls. Proceed-
loads. Proceedings of ASCE Specialty Conference on Lateral Stresses ings of ASCE, Symposium on Earth Reinforcement, Pittsburgh, PA,
in the Ground and Design of Earth-Retaining Structures. ASCE 1978. p. 664–84.
Special Publication No. 25, 1970. p. 103–47. [26] Fairless GJ. Seismic performance of reinforced-earth walls. Research
[5] Elgamal AW, Alampalli S, Laak P. Forced vibration of full-scale Report 89-8, Department of Civil Engineering, University of Canter-
wall-backfill system. Proceedings of ASCE, Journal of Geotechnical bury, Christchurch, New Zealand, 1989.
Engineering 1996;122(10):849–58. [27] Zhao C, Steven GP. Asymptotic formulae for correcting finite element
[6] Veletsos AS, Younan AH. Dynamic soil pressures on vertical walls— predicted natural frequencies of gravity and embankment dams. Soil
state of the art. Proceedings of the 3rd International Conference on Dynamics and Earthquake Engineering 1996;15:161–9.
K. Hatami, R.J. Bathurst / Soil Dynamics and Earthquake Engineering 19 (2000) 137–157 157

[28] Itasca Consulting Group. FLAC, Fast Lagrangian Analysis of Conti- [36] Tatsuoka F, Koseki J, Tateyama M. Performance of geogrid-rein-
nua, Version 3.40, Itasca Consulting Group Inc., Minneapolis, USA, forced soil retaining walls during the great Hanshin-Awaji earth-
1998. quake, 17 January 1995, Proceedings of the 1st International
[29] Bathurst RJ, Hatami K. Earthquake response analysis of reinforced- Conference on Earthquake Geotechnical Engineering, Tokyo,
soil retaining walls using FLAC. International FLAC Symposium on Japan, vol. I, 1995. p. 55–62.
Numerical Modeling in Geomechanics, Minneapolis, Minnesota, [37] Newmark NM. Effect of earthquakes on dams and embankments.
1999. p. 407–15. Geotechnique 1965;15(2):139–59.
[30] Murata O, Tateyama M, Tatsuoka F. Shaking table tests on a large [38] Bathurst RJ, Cai Z, Simac MR. Seismic performance charts for
geosynthetic-reinforced soil retaining wall model. In: Tatsuoka F, geosynthetic-reinforced segmental retaining walls. Proceedings of
Leshchinsky D, editors. Proceedings of Seiken Symposium No. 11, Geosynthetics 97, Long Beach, California, vol. II, 1997. p. 1001–14.
Recent Case Histories of Permanent Geosynthetic-Reinforced Soil [39] Bathurst RJ. NCMA segmental retaining wall seismic design proce-
Retaining Walls, Balkema, Tokyo, Japan, 1992. p. 259–64. dure—supplement to design manual for segmental retaining walls.
[31] Sakaguchi M, Yamada K, Tanaka M. Prediction of deformation of Herdon, VA: National Concrete Masonry Association, 1998.
geotextile reinforced walls subjected to earthquakes. Proceedings of [40] Cai Z, Bathurst RJ. Seismic-induced permanent displacement of
the 5th International Conference on Geotextiles, Geomembranes and geosynthetic-reinforced segmental retaining walls. Canadian
Related Products, Singapore, vol. I, 1994. p. 521–4. Geotechnical Journal 1996;33:937–55.
[32] Tatsuoka F, Tateyama, M, Murata O, Tamura Y. Geosynthetic-rein- [41] Arias A, Sanchez-Sesma FJ, Ovando-Shelley EA. Simplified elastic
forced soil retaining wall structures with short reinforcement and a model for seismic analysis of earth-retaining structures with limited
rigid facing-closure. In: Tatsuoka F, Leshchinsky D, editors. Proceed- displacements. Proceedings of the 1st International Conference on
ings of Seiken Symposium No. 11, Recent Case Histories of Perma- Recent Advances in Geotechnical Earthquake Engineering and Soil
nent Geosynthetic-Reinforced Soil Retaining Walls, Balkema, Tokyo, Dynamics, St. Louis, Missouri, 1981. p. 253–40.
Japan, 1992. p. 323–42. [42] Duncan JM, Byrne P, Wong KS, Mabry P. Strength, stress-strain and
[33] Tatsuoka F, Koseki J, Tateyama M, Munaf Y, Horii K. Seismic stabi- bulk modulus parameters for finite element analyses of stresses and
lity against high seismic loads on geosynthetic-reinforced soil retain-
movements in soil masses. Report No. UCB/GT/80-01, University of
ing structures. Keynote Lecture. Proceedings of the 6th International
California, Berkeley, California, 1980.
Conference on Geosynthetics, Atlanta, GA, vol. I, 1998. p. 103–42.
[43] Arias A. A measure of earthquake intensity. In: Hansen R, editor.
[34] Earthquake Engineering Research Centre. Large-scale shaking tests
Seismic Design of Nuclear Power Plants, Cambridge: MIT Press,,
of geotechnical structures. In: Taylor, CA editor. European Consor-
1969. p. 438–83.
tium of Earthquake Shaking Tables Prenormative Research in
[44] Housner GW. Intensity of earthquake ground shaking near the causa-
Support of Eurocode 8. University of Bristol, UK, 1997.
tive fault. Proceedings of the 3rd World Conference on Earthquake
[35] FHWA. Mechanically stabilized earth walls and reinforced soil slopes
Engineering. Wellington, New Zealand, vol. I, 1965. p. III’94–
design and construction guidelines. Elias V, Christopher BR, editors.
III’115.
Federal Highway Administration (FHWA) Demonstration Project 82,
Washington, DC, 1996.

You might also like