You are on page 1of 24

Computers and Geotechnics 17 (1995) 523-546

0 1995 Elsevier Science Limited


Printed in Great Britain. All rights reserved
0266-352X/951$9.50
ELSFVIER

SEISMIC RESPONSE ANALYSIS OF GEOSYNTHETIC REINFORCED SOIL

SEGMENTAL RETAINING WALLS BY FINITE ELEMENT METHOD

Zhenqi CAI and Richard J. BATHURST

Department of Civil Engineering, Royal Military College of Canada, Kingston, Ontario,

Canada K7K 5LO

ABSTRACT
The paper presents the results of a finite element analysis of the dynamic response of a
geosynthetic reinforced soil retaining wall that is constructed with dry-stacked modular
concrete blocks as the facia system. In the finite element model, the cyclic shear behavior of
the backfill soil is described by a hyperbolic stress-strain relationship with Masing hysteretic
unload-reload behavior. The reinforcement material is modelled using a similar hysteretic
model which takes into account the measured response of cyclic load-extension tests
performed on unconfined geogrid specimens in the laboratory. Interface shear between wall
components is simulated using slip elements. The results of finite element analyses giving the
seismic response of a typical geogrid reinforced segmental retaining wall subjected to
prescribed acceleration records are presented. The results of analyses highlight the influence
of dynamic loading on: (1) wall displacement; (2) cumulative interface shear force and
displacement between facing units; (3) tensile forces developed in the reinforcement and; (4)
acceleration response over the height of the wall. A number of implications to the design of
these structures are identified based on the results of these simulations.

INTRODUCTION

The use of segmental retainingwalls that consist of a column of discrete concrete block units
as the facia system together with geosynthetics to anchor the facia and reinforce the retained
soils has gained popularity in Europe and North America in recent years [ 11. A comprehensive
methodology to analyze and design these systems under static conditions has been recently
introduced by the National Concrete Masonry Association (NCMA) in North America [2,3].
The economic benefits of geosynthetic reinforced soil walls over conventional concrete gravity
walls and mechanically stabilized soil retaining walls that use inextensible (steel)
reinforcements has been demonstrated in a number of publications e.g. [4-61. The essential
components of a segmental retaining wall are illustrated in Figure 1.
Field measurement results have shown that limit-equilibrium based methods of analysis are
likely conservative for reinforced segmental walls under static loading conditions [7,8].
523
524

Modular Concrete
FACING UNIT
REINFORCED SOIL ZONE
I

6.0m

0.2m _. I

i1 i
I

fZ i
7
-ww
0.6m
h------- 4.3m W

FIGURE 1 Geogrid reinforced soil wall with discrete modular concrete facing units
(after Bathurst et al. 1993 [7])

However, the stability of these systems under dynamic loads induced during seismic events has
not been adequately addressed. Questions related to the performance of the discrete facia
system and the connections between the facia units and geosynthetic reinforcement layers that
are the distinguishing construction features of these walls have been raised [9]. Nevertheless,
the resiliency of geosynthetic soil reinforced walls and slopes (including a number of segmental
walls) under seismic loading during the recent Loma Prieta earthquake of 1989 and the
Northridge earthquake of 1994 has been qualitatively demonstrated [lO,ll].

The most common approaches for seismic stability analyses of geosynthetic reinforced soil
retaining walls are based on pseudo-static limit-equilibrium methods. These methods
calculate dynamic earth pressures using the Mononobe-Okabe method or a modified two-part
wedge method, which are essentially the same approach that has been used for many years for
the stability analysis of conventional gravity retaining wall structures, e.g. [12,13]. Current
pseudo-static seismic design methods for reinforced soil walls are a logical extension of the
Coulomb wedge method used for the analysis of reinforced walls under static loading [14,15].
However, experience with gravity wall structures has shown that pseudo-static methods are
excessively conservative e.g. [15]. In order to give reasonable predictions of wall stability,
empirical reductions of dynamic forces have often been employed [14,16]. Furthermore, large
body displacements observed for many conventional gravity structures which have
525

nevertheless remained intact cannot be predicted by limit-equilibrium based approaches [l].


“Displacement methods” that treat the failed soil mass and gravity wall structure as separate
rigid bodies have been proposed to overcome the dissatisfaction with limit-equilibrium based
methods applied to conventional gravity structures [13,17,18]. Vrymoed [19] proposed a
seismic design method for reinforced soil walls that considers the pseudo-static equilibrium
of the wall as a rigid block and the retained soil as a rigid wedge. Newmark’s method [20] was
then applied to estimate permanent wall displacement, which was considered to occur by
sliding at the base of the wall and/or by pullout of the reinforcing elements causing an outward
tilting of the wall face. This is similar to the methodology developed by Richards and Elms
[13] for gravity retaining walls. However, this method can be expected to introduce further
complications when complex materials such as rate dependent polymeric reinforcements are
considered together with discrete facing elements.

Properly formulated finite element models (FEM) offer the attraction of being able to
simulate the deformation response of reinforced soil structures subjected to seismic loading.
The deformation response is arguably the most important performance feature of these
structures under seismic loading.

The finite element method has been used to investigate the static response of a number of
geosynthetic reinforced soil structures [21-241. However, the use of FEM to investigate the
dynamic response of these structures has been limited. Recently, Bachus et al. [25] carried out
a parametric study of reinforced earth walls using the program DYNA3D. Segrestin and
Bastick [26] reported results using the FEM program SUPERFLUSH to analyze the response
of two full-scale reinforced earth walls that used large articulated concrete panel facings and
inextensible steel strips as the soil reinforcement.

A research program is currently underway at the Royal Military College (RMC) of Canada
that is focused on the study of the response of geosynthetic reinforced soil structures under
dynamic excitation due to earthquake and blast loadings. A2-D dynamic finite element code,
TARA-3, originally developed at the University of British Columbia [27], has been modified
and extended to account for the effects of extensible reinforcing elements subjected to cyclic
loading. Acyclic non-linear hysteretic constitutive relationship for the extensible geosynthetic
material has been proposed and incorporated into the program as described later in this paper.
The original program has been demonstrated to give excellent agreement with the measured
dynamic response of an instrumented full-scale reinforced earth wall that used large
articulated concrete panels anchored with steel strips [28]. The modified version of the
program has been used to simulate the dynamic response of a reinforced soil slope and a
soil-retaining wall constructed with extensible polymeric reinforcement [29,30].

The present paper is focused on the seismic response of a special class of geosynthetic
reinforced soil retaining wall systems that use dry-stacked modular concrete blocks as the
facia. These systems are unique in that the facia blocks are discrete and special attention must
be paid to proper modelling of block to block and block to reinforcement interface behavior
526

and the connection between the reinforcing elements and the facia. The paper also outlines
the constitutive models used to represent non-linear cyclic load response of the soils and
reinforcement elements.

FINITE ELEMENT SIMULATION

Seismic response analyses of a typical geosynthetic reinforced soil retaining wall


constructed with modular concrete facing units and a geogrid reinforcement were carried out.
The wall was designed for static loading using a conventional “tie-back wedge” method of
analysis together with Coulomb earth pressure theory and factors of safety recommended by
the NCMA for routine structures [2,3].
The finite element mesh used to carry out the simulations includes 650 elements and 657
nodes and is illustrated in Figure 2. The wall is 3.2 m high and comprises 16 concrete masonry
units, a uniform granular backfill and five layers of high density polypropylene (HDPE)
geogrid reinforcement extending 2.5 m into the backfill soil. The modelled length of the
backfill soil extends a distance of 14 m. The soil mass is discretized into 4-noded isoparametric
elements with the exception of a few transition triangular elements as shown in the figure.
Each facing unit is modelled as a single quadrilateral element with linear elastic properties.
The geogrid layers are represented by tension-only reinforcement elements. Interface slip

slip elements:

block/soil block/geogrid block/block soil/geogrid soil/base block/base

FIGURE 2 Component materials and finite element mesh for simulation of example
segmental retaining wall
527

behavior at any interface is modelled by slip elements identified in the figure. For simplicity,
the foundation soil is assumed to be rigid and the right side of the retained soil mass is assumed
to be an elastic energy transmitting boundary. The base reference acceleration-time history
used is a scaled El-Centro 1940 earthquake record [311 and is shown in Figure 3. Spectrum
analysis of the input acceleration record gives a dominant frequency range of 0.5 to 2 Hz. The
constitutive models adopted in the current study and selected parameters are briefly described
in the following sections.

‘---0.219
-.25
0 5 10 15 20 25 30
time (s)
FIGURE 3 Time history of base reference input acceleration [31]

CONSTITIJTIVE MODELS FOR COMPONENT MATERIALS

Soil model
The failure criterion for the soil used in TARA- is the Mohr-Coulomb shear failure
criterion. The shear behavior of granular soils under cyclic loading is modelled using a
non-linear and hysteretic constitutive relation that follows the Masing rule during unloading
and reloading [27]. The model is illustrated in Figure 4. The relation between shear stress t
and shear strain y for the initial loading phase (along the stress-strain cap) is described by the
following equation:

G max Y
t = f (y) = (1)
1+ G,,x/ r,, 1~1
[ ( 1 I
where G,, is the initial shear modulus and tmax is the maximum shear stress. The Masing
stress-strain equation for the unloading or reloading stage is given by:

‘G -
- zr = Gmw (Y - or) / 2
2
1 + &x+x / 2r,,$y - Y~I]
[ (
where (yr ,zr ) represents the stress state at which the shear stress reverses direction.
528

t z
t _____---------- tmax
t ____----------~ Tmax
stress-strain cap

-
f
(a2- 1
Y Yr

Y
y unload
a) stress-strain cap b) unload-reload

FIGURE 4 Non-linear hysteretic stress-strain model for granular soil

Assuming isotropic soil properties and plane strain conditions, the constitutive relationship
between incremental stresses, {Au)= {Au,, Aq, At,>, and incremental strains, {AE} = {AE,,
AE,,, Ay,}, is given as follows:

{ho) = [Dl {AC) (3)


where [D] is the tangent elasticity matrix. In the present analysis, [D] is formulated using the
tangent shear modulus (Gt) and the tangent bulk modulus (B,):

PI = BdQ,l + WQ,l (4)


in which [Ql] and [Q2] are two constant matrices given by:

[al]= [b ; i] ,,,]=Ez -+3 fl (5)

The tangent shear modulus (GJ at a point on the stress-strain cap is given by:

Gd!?= G max
t
du 2 (f-5)

At a point on an unloading or reloading curve Gt is given by:

G
G, =
(7)

[
I +
(G,, / 21:&y - yrl]’

The initial dynamic shear modulus of the soil (G,,) is calculated using an empirical
relationship proposed by Seed and Idriss [32]:

G max = 21.7 K2,,,= P, (p)” (8)


a
529

in which o,,, is the mean normal stress, Pa is the atmospheric pressure and Kzrnax is the shear
modulus constant.
The tangent bulk modulus, Bt , is expressed in the following form:

B, = K, P, (p)” (9)
a
in which Kt, is the bulk modulus constant and n is the bulk modulus exponent. Methods to
estimate constants KzmawrKt, and n can be found in [32,33].
Reinforcement model
The load-strain cap/hysteretic unload-reload model adopted for the soil materials has been
modified for polymeric reinforcement materials [30] and is illustrated in Figure 5. The
relationship between axial load and tensile strain for the load-strain cap is expressed by:
Di sa
F = [ 1 + ( Di / F, ) IEaI I (10)

where F is the axial tensile load per unit width of specimen (e.g. kN/m), E, is the axial strain,
Di is the initial modulus and F, is an extrapolated asymptotic ultimate strength of the
reinforcement material.
During an unload-reload cycle, the reinforcement model is assumed to follow the Masing
rule. The equation for the unloading curve from point A(+, Fr) or, for the reloading curve
from point B at which the load reverses direction is thus given by:

-F - F, = D,, (&a - Er)/ 2


01)
2 [ 1 + ( D,, / 2 F, ) 1% - sr] I
Here, D,, is the unload modulus defined in terms of the initial load modulus, D,r=kDi, with
k equal to a constant.

load-strain cap
Di sa
= t 1 + ( Di / Pm ) lsal I

I
J F-F, = Dur(&a or)/ 2

1
-

B
2 [ 1 + ( Dur / 2 Fm 1 Isa - ELI I

axial strain E,

FIGURE 5 Cyclic unload-reload model for polymeric reinforcement


530

Slin element model

The shear transfer between any two contacting materials such as, reinforcement and soil,
reinforcement and concrete block, or block to block, is an important mechanism that
influences the response of the entire reinforced wall structure during seismic loading.
Interfaces are represented by slip elements. The interface slip model adopted in the current
study is illustrated in Figure 6 and follows that proposed by Goodman et al. [34]. The failure

= c, + ontan+,
t “yield

‘yield

relative displacement

FIGURE 6 Interface slip model

(yield) state of the slip element is assumed to obey the Mohr-Coulomb criterion:

‘yield
= cs + u,, tan &, where ryi& is the shear strength at which slip occurs for the first time;
c, is the apparent cohesion; on is the normal stress; and Qs is the interface angle of friction at
the yield state. When the applied shear stress exceeds the yield strength, the shear stiffness
of the slip element is reduced to a fraction (a) of the original value and slip is initiated. In the
normal direction, the stress o, is assumed to vary linearly with the average relative nodal
displacement. Separation of the contact interface is assumed to occur when the normal stress
un is tensile.

A detailed description of the theory and solution of the dynamic incremental equations of
motion in program TARA- can be found in [27].

SELECTION OF MATERIAL PROPERTIES

Soil element nronerties

The selection of material parameters Kt,, Kzrnax and n used in Equations 8 and 9 to describe
the properties of the backfill soil was based on a typical granular soil [33]. These parameter
values together with the assumed soil friction angle and unit weight are given in Table 1. The
foundation soil is assumed to be rigid and the right side boundary of the retaining wall is an
531

elastic energy transmitting boundary as proposed by Lysmer and Kuhlemeyer [35], with an
assumed P wave velocity of 400 m/s and a S wave velocity of 200 m/s.

TABLE 1

Backfill soil properties


Kb K2max n @S YS
700 70 0.5 350 19 (kN/m3)

Reinforcement element properties


In-isolation cyclic load-controlled tests on typical polymeric geogrid reinforcement
materials have been reported by Bathurst and Cai [36]. The parameters for the reinforcement
model introduced earlier can be estimated from the results of this laboratory study. Figure 7
shows results of a multi-stage cyclic load-extension test on a HDPE geogrid specimen. The
figure illustrates that the load-strain cap is well-represented by a hyperbolic curve fitted to the
measured data and that hysteretic behavior during unload-reload cycles is significant above
0.5% strain. Figure 8 shows variation of the initial stiffness modulus Di with the frequency
of loading. The initial tangent modulus of the HDPE geogrid increases with an increase in
loading frequency. Thus, for a given seismic record, the values assigned to parameters Di , D,,
and F, should be based on the dominant frequency content of the motion when a HDPE
geogrid reinforcement is used. The data in Figure 8 also illustrates that stiffness modulus

70

freauencv 1 .O Hz
r

i
0 1 2 3 4 5 6 7 8
strain (%)
FIGURE 7 In-isolation cyclic load test on a HDPE geogrid
532

4000
stiffness value from
3500 index test at 10% strainlmin 7

//‘I L__w
range of test data

I III, I
0.10 1 .oo 10.00
frequency (Hz)

FIGURE 8 Initial stiffness modulus for HDPE geogrid under cyclic loading
(after Bathurst and Cai [36])

values obtained from standard constant rate of extension (index) tests may underestimate the
stiffness of HDPE geogrids for rates of loading anticipated during a seismic event.

The parameter values inferred from the cyclic load test carried out at 1 Hz were used in the
simulations and are given in Table 2. The choice of 1 Hz falls within the predominant excitation
frequency of the acceleration record (Figure 3) used.

TABLE 2
HDPE geogrid reinforcement properties at a loading frequency of 1 Hz

Di F, D,r
3080 kN/m 125 kN/m 1.4

The slip element properties used in the current study are given in Table 3 for the six
interface conditions identified in Figure 2. The shear stiffness k, of each type of interface
element is assumed to vary proportionally with the square root of the normal stress. The values
in Table 3 for block to block shear transfer were interpreted from a direct shear test on typical
modular block units [37]. An example experimental result for the actual modular block units
used in the numerical simulations is illustrated in Figure 9. It should be noted that the shear
transfer between modular block units is influenced by interface geometry such as shear keys
and the presence of mechanical connectors for some systems [l-3]. The test results used here
are for a modular block system with no shear keys or mechanical connectors and hence
represent a relatively poor system from a shear capacity point of view. The parameters used
533

for interface shear between block and reinforcement were interpreted from results of a
connection test with a HDPE geogrid and a similar block system [38] and are illustrated in
Figure 10. The results of pullout tests carried out at RMC using a sand and a HDPE geogrid
[39] were used to infer the soil/reinforcement interface properties given in Table 3. The HDPE
geogrid in the above referenced pullout and connection tests was the same geogrid assumed
in the current study.

The block to base interface and soil to base interface stiffness values were arbitrarily
doubled in magnitude from those used for the block to block interface. This was done to
account for the effects of embedment depth (i.e. 0.5 m is recommended for typical segmental

TABLE 3
Parameters used for interface elements

block/soil1 block/grid* block/block3 soil/grid4 soil/base5 block/base6

k, 9000 9600 9000 9600 18000 18000


6 6
k, lo6 106 106 106
0.05 01.;5 0.05 0.25 0.25 01025
;s 23O 250 38“ 350 35O 38O

k, shear stiffness (kN/m3) a yield shear stiffness factor


k, normal stiffness (kN/m3) QS friction angle at yield
1. a,=42 kPa (assumed) 2. a,,=79 kPa (Figure 10 [38])
3. a,,=42 kPa (Figure 9 [37]) 4. a,=69 kPa [39]
5. a,=42 kPa (assumed) 6. a,=42 kPa (assumed)

an = 42 kPa f a,=79 kPa

40 I

*
0 5 10 15 20 25 30 0 5 10 15 20
relative displacement (mm) relative displacement (mm)

FIGURE 9 Example direct shear test FIGURE 10 Example connection test


for block to block interface for geosynthetic to block
interface
retaining walls [2]) and the assumed incompressible foundation, both ofwhich result in a stiffer
wallibase interface.
Related work reported by Lahlaf and Yegian [40] on interface shear behavior of
geosynthetic layers during shake table tests has shown that the difference between the dynamic
friction angle and the static friction angle between geosynthetic layers is negligible. This gives
some support to the assumption made in this study that interface friction angles from the
results of static shear tests can be used to represent dynamic conditions.
Facing element uronerties
The properties of the modular facing blocks were assigned the following values: elastic
modulus, EC=26 GPa; Poisson’s ratio, w=O.2; and unit weight, yc=22 kN/m3. Quadrilateral
elements were used to model these units and the block dimensions were chosen to match the
specimen size reported for the direct interface shear tests (300 mm x 200 mm).

COMPUTATIONAL DETAILS

The results from a static load (end of construction) analysis of the example wall shown in
Figure 2 were used to provide initial stress conditions and moduli properties for subsequent
seismic simulations. The construction of the wall was assumed to be completed in 16
incremental layers to match the number of facing blocks. Each layer construction was
simulated using 20 load steps. Load shedding was accounted for in the backfill soil during each
load step. The modulus parameters at the final stage of the construction were used as initial
values for the dynamic analyses. Numerical integration of the system equations of motion
during seismic excitation was implemented using the Wilson-f3 method (8=1.4). A maximum
number of 6 iterations of calculation were carried out to account for load reversals during each
time increment. The stiffness and damping matrices in the system equations of motion were
updated at each loading increment as a function of the current tangent moduli of the soil. A
more complete description of computational details of TARA- can be found in [27,28].
Loading history
The horizontal acceleration-time record in Figure 3 was scaled with respect to the peak
acceleration to give two base input acceleration-time histories for the example reinforced soil
wall structure. The peak accelerations were 0.125g and 0.25g and each acceleration history
was applied for the same 30 second duration shown in Figure 3.

SIMULATION RESULTS

Wall lateral displacement


The displacement-time histories at selected locations along the wall face due to the
excitation with a peak acceleration of 0.25g are shown in Figure 11. The data shows that
displacements are cumulative over the duration of excitation and are permanent at the end of
535

lent

0.259
4 *
peak acceleration

97

& 97 Layer 1
J-

, , , , , , , , , , , , , , , , ,
2 4 6 8 10 12 14 16 18 20 22 24 26 28 :
elapsed time (seconds)
FIGURE 11 Displacement-time histories at selected locations

excitation. This cumulative pattern of wall lateral displacement is similar to experimental


observations reported for other types of reinforced soil structures that were subjected to
similar acceleration histories [41,42] (peak acceleration scaled differently). The permanent
lateral deformation of the wail front face at both excitation levels is given in Figure 12. The
datum used is the static wall position. Figure 12 shows that a large portion of the
seismic-induced permanent displacements occurred within approximately the bottom l/3 to
l/2 wall height above the base. A similar pattern has been observed from the results of
instrumented segmental retaining walls under static loading [7,8]. Relative shear
536

peak acceleration
3.2T- ~_~._._
0.1259

2.8.

reinforcement

2.4-

2.2.

2.0-

E 1.8- datum =
static wall
E kt position
9 ‘6
al
z 1.4
z
1.2-

l.O-

0.8-

0.6

0.4-

0.2.

-25 -20 -15 -10


displacement
-5
(mm)
u 0
of&,* ,

FIGURE 12 Lateral displacement profiles along wall facing at end of


excitation history

displacements can be observed at all block to block and block to geogrid interfaces over the
height of the wall. This suggests that these interfaces may be a potential cause of wall collapse
under seismic loading.

Wall interface shear behavior

An example of cumulative shear force at the layer 1 block to reinforcement interface is


illustrated in Figure 13. As expected from the shear-displacement interface model adopted
in the simulation, the magnitude of shear developed across the interface is also cumulative and
is dependent on both the duration and magnitude of the base excitation input. A long duration
earthquake with relatively lower peak acceleration can lead to the same cumulative shear
static

I
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
elapsed time (seconds)

FIGURE 13 Time history of shear force along block to geogrid interface

force as an excitation input of shorter duration but with greater peak acceleration. This
cumulative pattern cannot be predicted by pseudo-static methods.

Figure 14 compares the interface yield shear capacity available along the height of the wall
with the computed peak shear forces developed under the two levels of input acceleration and
under static load conditions. Also shown on the figure are the yield shear strength envelopes
for block to block and block to reinforcement based only on the self-weight of the column of
blocks above the interface. Wall-soil interface friction and lateral earth pressure on the back
of the wall result in additional normal stresses being carried by the column of wall units and
this accounts for the greater shear capacity calculated from the results of the FEM analysis
than from the self-weight alone. The results presented in Figure 14 also illustrate that the
block-reinforcement interface generates the greatest local shear transfer. The peak shear
forces induced by the 0.25g seismic record at layer 1,2 and 4 exceed the corresponding yield
shear capacities and result in the relatively large slip displacements predicted at these levels
(Figure 12). Hence the choice of interface shear values where a geosynthetic inclusion is
present is particularly important for the seismic design of segmental geosynthetic reinforced
soil retaining walls. However, it should be remembered that the selected interface shear
parameters in the present study correspond to segmental units that have weak shear interface
properties. Results of interface shear testing of other modular block systems have shown that
modular blocks formed with shear keys can develop high interface shear stiffness and strength
both with and without the presence of a reinforcement inclusion [1,37].
538

Static 0 block-reinforcement

IH
3. lp o block-block
2 8 2_0?5a
3 !-=Jg_ block-reinforcement
2. capacity based on

static
shear
strength
envelope
(FEW

5 10 15 20 25
peak shear force (kN/m)
FIGURE 14 Interface shear at end of excitation history and under static load

Reinforcement forces

The tensile force distributions along each reinforcement layer at the end of the static
loading (end of construction) and at the end of the two excitation records are shown in Figure
15. The trend towards reduced tensile force in the reinforcement layers close to the facia units
is consistent with large-scale model tests carried out at RMC that used flexible facia systems
[43] and the results of instrumented field structures that used modular block units [7,8]. The
same pattern of tensile force is observed after dynamic excitation but tensile load levels are
higher as expected. An exception to the trend of increased tensile load with increases in
acceleration is reinforcement layer 3. It is possible that this layer is located close to a pivot
node of the dominant eigenmode of the wall system. In addition, the pattern of peak load for
layer 4 in Figure 15 shows a shift towards the front of the wall. The distribution of the maximum
tensile load in each reinforcement layer with layer elevation is presented in Figure 16. This
539

& - static 1 I

g a) --- 0.1259 Layer 5


---- 0.2509 __-----_
4-
--=__ --__-_--- -----_‘=-_
2- w>_
-_e __--_-z= - - \
or, I1 I /I,IIII 1 l,,,,,,,\--T
8
g b) Layer 4
_/--- ----___ --__-- ___----_
4- -_
,_-- _____------ ------___
:;:
2-/’ _--
07’,S,S 7lIIIIII 1 ,CII,,,/II,

cl Layer 3

8
g e) Layer 1

0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
distance from center of facing column (m)

FIGURE 15 Tensile force in reinforcement layers

plot illustrates that the large forces in layers 1 and 4 are likely the result of amplification of
shear displacements associated with the predominant eigenmodes of the composite structure.

Superimposed on Figure 16 are the static reinforcement forces calculated using Coulomb
active pressure theory and the total pseudo-static forces using the Mononobe-Okabe method
with a peak acceleration of 0.25g. Both methods yield conservative results as compared to the
corresponding FEM analyses.

Figure 17 shows the time history of tensile load in two reinforcement layers. The comments
made earlier regarding the accumulation of wall lateral displacements and shear forces at
block to block and block to reinforcement interfaces also apply to tensile loads developed in
the reinforcement elements. For example, a long duration seismic event at relatively low peak
acceleration can generate the same tensile load as a higher peak acceleration record of shorter
duration. Pseudo-static methods cannot account for this phenomenon. However, it should be
noted that stress-relaxation can occur over time in polymeric reinforcement layers which was
540

static (Coulomb) Mononobe-Okabe

2.4

0.6

0.2
static 0.125g 0.2509
0 I I I I I I
0 1 2 3 4 5 6 7 8 9
peak tensile force in reinforcement (kN/m)

FIGURE 16 Peak reinforcement forces

static

0 5 10 15 20 25 30 0 5 10 15 20 25 30
elapsed time (seconds) elapsed time (seconds)

FIGURE 17 Example reinforcement tensile force-time response


541

not accounted for in the current simulations. Hence, the seismic-induced forces in the
reinforcement layers may not persist over time once a static load condition is re-established,
Distribution of horizontal accelerations
Figure 18 shows the horizontal acceleration predicted at different locations in the
unreinforced soil mass, reinforced soil mass and the facia column, along the mid-wall height
level, for a duration of 15 seconds. The data illustrates that peak accelerations occur at the
same time across the entire structure. Data extracted from similar locations along the l/3 and
2/3 wall height elevations gave the same result. These observations have important
implications to pseudo-static methods of design in which it is often assumed that peak dynamic
lateral earth pressures and wall inertial forces due to seismic excitation do not occur
simultaneously and therefore the dynamic earth forces are usually reduced. The dynamic
earth pressure is typically reduced by 40% applied to an arbitrary selected portion of the
reinforced soil mass (i.e. 0.5ayH2 where a is the horizontal acceleration ratio) [14,16].
Figure 19 shows the peak acceleration profile along the height of the wall within the facia
column, the reinforced soil zone, and the retained soil mass. The positive acceleration acts
towards the soil and creates an inertial force that adds to the static driving force (active) on
the wall facing, while the negative acceleration does the opposite. The data shows that the
variation in the peak acceleration distribution along the height of the wall is small. This
supports the use of a single acceleration factor in pseudo-static methods. It is also seen that
the positive peak acceleration is generally smaller than the ground input acceleration peak
(except at top of wall facing) due to system damping while the negative peak is amplified with
height above the toe of the wall.

CONCLUSIONS AND IMPLICATIONS TO DESIGN

The application of a dynamic finite element program (TARA-3) to investigate the seismic
response of a typical geogrid reinforced modular block wall has been reported. The results
of analyses highlight the influence of dynamic loading on: (1) wall displacement; (2)
cumulative interface shear force and displacement between facing units; (3) tensile forces
developed in the reinforcement and; (4) acceleration response over the height of the wall.
None of these response features can be examined by pseudo-static methods of analysis.
The following observations have important implications to the seismic design of reinforced
segmental retaining wall structures:

1. Relative displacement and interface shear force between modular units increased with
the duration and magnitude of base excitation during simulated seismic events.
Relative displacements and shear forces were greatest at interfaces where a
geosynthetic was present and at some locations the shear capacity between modular
units was exceeded. Hence an accurate estimate of interface shear properties is
542

reinforced zone retained zone

0.3
0.2- .2329,2.44s, node 318 (block)
,149g.S 16s
.118g,12.06sli _* d h
0.1
t Ah ,n

‘1
I1, v
-0.2. I -.17og,5.3 8s Y -._ ._ ._
- -.217g,12.4us

-"'30 1 2
.277g,2.6Os. 3 4 5 6 7 8 9 10 11 12 13 14 15

$ 0.3
8
Cfj 0.2
0.1
-0.0
-0.1
-0.2

-“.30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

-“‘30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
elapsed time (s)

FIGURE 18 Horizontal acceleration response at different locations


along the mid-height wall level
- - - - - retained soil
--- reinforced soil
- facing blocks I’
Iv
1: I
\ I
I

i!I

-0.35 -0.25 -0.15 -0.05 0.05 0.15 0.25 0


peak acceleration (g)

FIGURE 19 Peak acceleration distribution profile along height of wall

particularly important for the seismic design of segmental geosynthetic reinforced soil
retaining walls.

Dynamic tensile forces in reinforcement layers were also observed to be cumulative


over the duration of the simulated seismic events. Peak tensile forces were observed
to increase with the magnitude of peak base acceleration but relative increases in
tensile forces were not uniform through the height of the structure.

Reinforcement forces calculated using pseudo-static earth pressures (M-O method)


were consistently greater than the forces calculated by the dynamic FEM analyses.
Reinforcement forces resulting from static load conditions and calculated using the
Coulomb method were also greater than the FEM results.

The horizontal accelerations predicted at different locations in the unreinforced soil


mass, reinforced soil mass and the facia column showed that peak accelerations
occurred at the same time across the wall system. Results also showed that variations
in the peak acceleration distribution along the height of the wall were small. This gives
544

some support to the conventional practice of using a single acceleration factor in


pseudo-static methods of analysis.

ACKNOWLEDGEMENTS

The funding for the work reported in the paper was provided by the Department of National
Defence (Canada) through an Academic Research Program (ARP) grant and a research
contract from Director of Architecture (D-ARCH).

REFERENCES

1. Bathurst, R.J. and Simac, M., Geosyntheticreinforced segmental retainingwall structures


in North America, Keynote paper, Proceedings of 5th International Conference on
Geotextiles, Geomembranes and related Products, Preprint: Special Lectures and Keynote
Lectures, Singapore, (1994) 3 l-54.
2. Simac, M., Bathurst, R.J., Berg, R.R. and Lothspeich, S.E., Design manualforsegmental
retaining walls, National Concrete Masonry Association, 2302 Horse Pen Road, Herndon,
Virginia, USA, (1993) 336~.
3. Bathurst, R.J., Simac, M.R. and Berg, R.R., Review of the NCMA segmental retaining
wall design manual for geosynthetic reinforced structures, Transport&ion Research Record
1414, (1993) 6-25.
4. Crowe, R.E., Bathurst, R.J. and Alston, C., Design and construction of a road
embankment using geosynthetics, Proceedings of 42’nd Canadian Geotechnical
Conference, (1989) 266-271.
5. Simac, M.R., Bathurst, R.J. and Goodrun, R.A., Design and analysis of three reinforced soil
retaining walls, Proceedings of Geosynthetics’91, Atlanta, GA, Vol.2, (1991) 781-789.
6. Hill, J.J. and Berg, R.R., Use of segmental wall system by Minnesota Department of
Transportation, Transpotiation Research Record 1414, (1993) l-5.
7. Bathurst, R.J., Simac, M.R., Christopher, B.R. and Bonczkiewicz, C., A database of
results from a geosynthetic reinforced modular block soil retaining wall, Proceedings of
Soil Reinforcement: Full Scale Experiments of the 80’s, ISSMFEiENPC, Paris, France,
(1993) 341-365.
8. Simac, M.R., Christopher, B.R. and Bonczkiewicz, C., Instrumented field performance of a 6m
geogrid wall, Proceedings of4th International Conference on Geotextiles, Geomembranes and
related Products, The Hague, Netherlands, Vol. 1, (1990) 53-59.
9. Allen, TM., Issues regarding design and specification of segmental block-faced
geosynthetic walls, Transportation Research Record 1414, (1993) 6-11.
10 Collin, J.G., Chouery-Curtis, V.E. and Berg, R.R., Field observations of reinforced soil
structures under seismic loading, Earth Reinforcement Practice, Ochiai, Hayashi & Otani
(eds), Balkema, Rotterdam, (1992) 223-228.
11. Sandri, D., Retaining walls stand up to the Northridge earthquake, Geotechnical Fabrics
Report, IFAI, St. Paul, MN, USA, Vol. 12, No. 4, (1994) 30- 31 (and personal
communication).
12. Seed, H.B. and Whitman, R.V., Design of earth retaining structures for dynamic loads,
ASCE Specialty Conference on Lateral Stresses in the Ground and Design ofEarth Retaining
Structures, Ithaca, NY, (1970) 103-147.
545

13. Richards, R. and Elms, D.G., Seismic behavior of gravity retainingwalls, J. Geotech. Eng.
Div., ASCE, Vol. 10.5, No. GT4, (1979) 449-464.
14. Christopher, B.R., Gill, S.A., Giroud, J-P, Juran, I., Schlosser, F., Mitchell, J.K. and
Dunnicliff, J., Reinforced soil structures, Volume I. Design and construction guidelines,
Federal Highways Administration, Washington USA, Report NO. FHWA-RD-89-043
(1989), 287~.
15. Bonaparte, R., Schmertmann, G.R. and Williams, N.D., Seismic design of slopes
reinforced with geogrids and geotextiles, Proceedings of 3rd International Conference on
Geotextiles, Geomembranes and related Products, Vienna, Austria, (1986) 273 -278.
16. Reinforced Earth Company, Reinforced Earth structures in seismic regions, Reinforced
Earth Company Report, (1991) llp.
17. Nadim, F. and Whitman, R.V, Seismically induced movement of retaining walls, J.
Geotech. Eng., ASCE, Vol. 109 No. 7, (1983) 91.5-931.
18. Lin, J-S and Whitman, R.V., Earthquake induced displacements of sliding blocks, J.
Geotech. Eng., ASCE, Vol. 112 No. 1, (1984) 44-59.
19. Vrymoed, J., Dynamic stability of soil-reinforced walls, Transportation Research Record
1242, (1989) 29-38.
20. Newmark, N.M., Effects of earthquakes on dams and embankments, Geotechnique 15, No.
2, (1965) 139-160.
21. Karpurapu, R. and Bathurst, R.J., Behavior of geosynthetic reinforced soil retainingwalls
using the finite element method, Computers and Geotechnics, accepted for publication
(1994).
22. Bathurst, R.J., Karpurapu, R. and Jarrett, PM., Finite element analysis of a geogrid
reinforced soil wall, Grouting, Soil Improvement and Geosynthetics, ASCE Geotechnical
Special Publication No. 30, Vol. 1, (1992) 1213-1224.
23. Ho, S.K. and Rowe, R.K., Finite element analysis of geosynthetic reinforced soil walls,
Proceedings of Geosynthetics ‘93, Vancouver, Vol. 1, (1993) 203-216.
24. Matsui, T. and San, EC., Prediction of two test walls by elastoplastic finite element
analysis, Int. Symp. on Geosynthetic Reinforced Soil Retaining Walls, Denver, Edited by
J.T.H. Wu, (1991) 143-151.
25. Bachus, R.C., Fragaszy, R.J., Jaber, M., Olem, K.L., Yuan, Z. and Jewell, R., Dynamic
response of reinforced soil systems, USAF Research Report FO 8635-90-C-0198, Civil
Engineering Laboratory, 139 Barnes Drive, Tyndall Air Fore Base, Florida, Vol. 1 and Vol.
2, June (1992) 457~.
26. Segrestin, l? and Bastick, M., Seismic design of Reinforced Earth retaining walls - the
contribution of finite element analysis, Proceedings Int. Geotech. Sym. on Theory and
Practice of Earth Reinforcement, Japan, (1988) 577-582.
27. Finn, W.D.L., Yogendrakimar, M. and Yoshida, N., TARA -3: a program to compute the
response of 2-D embankment and soil-sttucture interaction systems to seismic loading,
Department of Civil Engineering, University of British Columbia, Vancouver, Canada,
(1986).
28. Yogendrakumar, M., Bathurst, R.J. and Finn, W.D.L., Dynamic response analysis of a
reinforced soil retaining wall, J. Geotech. Eng., ASCE, Vol. 118, No. 8, (1992) 1158- 1167.
29. Yogendrakumar, M., Bathurst, R.J. and Finn, W.D.L., Response of reinforced soil slopes
to earthquake loadings, Proceedings 6th Can. Conf: on Earthquake Eng., Toronto, (1991)
445-452.
546

30. Yogendrakumar, M. and Bathurst, R.J., Numerical simulation of reinforced soil


structures during blast loads, Transportation Research Record 1336, TRB, (1992) l-8.
31. Romstad, K.M., personal communication (1993).
32. Seed, H.B. and Idriss, I.M., Soil moduli and dampingfactorfor dynamic response analysis,
Report EERC 70-10, Earthquake Research Engineering Center, University of
California, Berkeley, Dec., (1970).
33. Duncan, J.M., Byrne, I?, Wong, K. S. and Mabry, P, Strength, stress-strain and bulk
modulus parameters for finite element analyses of stresses and movements in soil masses,
Report No. UCB/GT/80-01, Department of Civil Engineering, University of California,
Berkeley, August, (1980) 70~.
34. Goodman, R.E., Taylor, R.L. and Brekke, T.L., A model for the mechanics ofjointed rock,
J. SMFE Div., ASCE, Vol. 94, (1968) 637-659.
35. Lysmer, J. and Kuhlemeyer, R.L., Finite dynamic model for infinite media, J. Eng. Mech.
Div., AXE, Vol. 95, EM4, (1969) 8.59-877.
36. Bathurst, R.J. and Cai, Z., In-isolation cyclic load-extension behavior of two geogrids,
Geosynthetics International, Vol. 1, No.1, (1994) 3-17.
37. Buttry, K., Wetzel, R. and McCullough, E., Testing of StoneWall landscape systems,
Department of Civil Engineering, University of Wisconsin at Platteville, April, (1991)
2op.
38. Masonry block wallconnection evaluation with selectgeogn’dproducts, Report prepared by
Geosyntec Consultants for Tensar Earth Technologies, Inc., Sept., (1991) 12~.
39. McLay, M.J., Pullout testing of geogrid reinforcement, MSc thesis, Department of Civil
Engineering, Royal Military College of Canada, April, (1993) 157~.
40. Lahlaf, A.M. and Yegian, M.K., Shaking table test for geosynthetic interfaces,
Proceedings of Geosynthetics’93, Vol. 2, Vanvouver, BC, (1993) 659-669.
41. Soon, D., Casey, J., Kutter, B.L. and Romstad, K.M., Dynamic centrifuge modelling of
sound walls supported on concrete cantilever and mechanically stabilized earth retaining
structures, Transportation Research Record 1336, TRB, (1992) 9-16.
42. Fairless, G.J., Seismicpedormance of reinforced earth walls, Research Report, Department
of Civil Engineering, University of Canterbury, New Zealand, Sept., (1989) 312~.
43. Bathurst, R.J., Jarrett, PM. and Benjamin, D.J.R.S., A database of results from an
incrementally constructed geogrid-reinforced soil wall test, Proceedings of Soil
Reinforcement: Full Scale Experiments of the 80’s, ISSMFE/ENPC, Paris, France,
November, (1993) 401-430.

Received 25 August 1994; revised version received 21 October 1994; accepted


28 October 1994

You might also like