Вы находитесь на странице: 1из 45

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/254560182

Pseudo-Static Seismic Analysis Of GeosyntheticReinforced Segmental Retaining Walls


ARTICLE in GEOSYNTHETICS INTERNATIONAL JANUARY 1995
Impact Factor: 1.68 DOI: 10.1680/gein.2.0037

CITATIONS

READS

37

336

2 AUTHORS, INCLUDING:
Richard J. Bathurst
Royal Military College of Canada
190 PUBLICATIONS 3,606 CITATIONS
SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Richard J. Bathurst


Retrieved on: 11 January 2016

Technical Paper by R.J. Bathurst and Z. Cai

PSEUDO-STATIC SEISMIC ANALYSIS OF


GEOSYNTHETIC-REINFORCED SEGMENTAL
RETAINING WALLS
ABSTRACT: The paper examines seismic stability analyses of geosynthetic-reinforced segmental retaining walls (modular block walls). Stability analyses are developed within the framework of a pseudo-static approach that gives factors of safety
against collapse mechanisms or rupture of component materials. The MononobeOkabe method is used to estimate dynamic earth pressures. Parametric analyses of
forces and factors of safety related to external, internal and facing failure modes for
walls constructed on competent foundations are presented. Shear interfaces between
facing units are considered as possible planes of failure in facing stability analyses. The
potential for local toppling of the facing column is also investigated. The results of analyses demonstrate that there is a limiting value of the horizontal seismic coefficient
above which the margin of safety against base sliding and overturning may be unacceptably low during a seismic event for segmental retaining walls designed to just satisfy
minimum factors of safety under static loading conditions. Pseudo-static seismic analyses of the performance of two geosynthetic-reinforced segmental walls during the
Northridge Earthquake in Los Angeles in 1994 are demonstrated to be consistent with
visual observation of tension cracks in the soil backfill. Limitations of pseudo-static
methods are discussed and recommendations for further research are made.
KEYWORDS: Pseudo-static analysis, Segmental retaining walls, Geosynthetic reinforcement, Seismic analysis, Modular block walls.
AUTHORS: R.J. Bathurst, Professor, and Z. Cai, Research Associate, Department of
Civil Engineering, Royal Military College of Canada, Kingston, Ontario, K7K 5L0,
Canada, Telephone: 1/613-541-6000 Ext. 6479, Telefax: 1/613-541-6599, E-mail:
bathurst@rmc.ca.
PUBLICATION: Geosynthetics International is published by the Industrial Fabrics
Association International, 345 Cedar St., Suite 800, St. Paul, MN 55101, USA,
Telephone: 1/612-222-2508, Telefax: 1/612-222-8215. Geosynthetics International is
registered under ISSN 1072-6349.
DATES: Original manuscript received 1 May 1995, revised manuscript received and
accepted 3 August 1995. Discussion open until 1 May 1996.
REFERENCE: Bathurst, R.J. and Cai, Z., 1995, Pseudo-Static Seismic Analysis of
Geosynthetic-Reinforced Segmental Retaining Walls, Geosynthetics International,
Vol. 2, No. 5, pp. 787-830.

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

787

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

INTRODUCTION

The use of segmental retaining walls that include dry-stacked concrete block units
as the facia system together with extensible sheets of polymeric materials (geosynthetics) that internally reinforce the retained soils and anchor the facia has gained wide popularity in North America (Bathurst and Simac 1994). These structures have also been
reported in Europe, Scandinavia and Australia in recent years (Cazzuffi and Rimoldi
1994; Gourc et al. 1990; Knutson 1990; Won 1994). An example of a reinforced segmental retaining wall structure is illustrated in Figure 1. The distinguishing feature of
these structures is the facing column that is constructed using mortarless modular concrete block units that are stacked to form a wall batter into the retained soils (typically
3 to 15_ from vertical). The economic benefits of these systems over conventional reinforced concrete gravity wall structures and mechanically stabilized soil retaining walls
that use inextensible (steel) reinforcement and select backfills have been demonstrated
in several of the references cited in an earlier paper by Bathurst and Simac (1994).
12kPa surcharge

Compacted
drainage fill

1.2m long geogrid type l

Masonry
concrete
facing units

Geogrid type ll

Geogrid type ll

5.5 m

20

Compacted drainage
fill with geotextile
Geogrid type ll

6.1 m
Compacted native soil
for reinforced soil zone
Geogrid type ll

Drainage collection pipe


gravity flow to outlet

0.3 m (typical)
Slope 1:6 (typical)

Excavation limits
4.4 m

Total geogrid embedment length (all layers)


Figure 1. Typical geosynthetic-reinforced soil segmental retaining wall cross-section
(after Simac et al. 1991).

788

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

Stability analyses for geosynthetic-reinforced soil walls under static loading conditions (including segmental retaining wall systems) involve separate calculations to establish factors of safety against external modes of failure and internal modes of failure
(Figure 2). External stability calculations consider the reinforced soil zone and the facing column as a monolithic gravity structure. The evaluation of factors of safety against
base sliding, overturning about the toe, and foundation bearing capacity is analytically
identical to that used for conventional gravity structures. Internal stability analyses for
geosynthetic-reinforced soil walls are carried out to ensure that the structural integrity
of the geosynthetic-reinforced soil mass is preserved with respect to reinforcement
over-stressing and pullout of geosynthetic reinforcement layers from the anchorage
zone.
A comprehensive design methodology has been recently proposed by the National
Concrete Masonry Association (NCMA) for the static analysis of segmental retaining
walls (Simac et al. 1993; Bathurst et al. 1993). The NCMA guidelines address potential

(a) base sliding

(b) overturning

(c) bearing capacity


(excessive settlement)

(d) tensile over-stress

(e) pullout

(f) internal sliding

(g) shear failure


(bulging)

(h) connection failure

(i) local overturning


(toppling)

Figure 2. Modes of failure: external (top row); internal (middle row); and facing (bottom
row) (Simac et al. 1993).

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

789

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

failure mechanisms not found in other geosynthetic-reinforced soil wall systems as illustrated in Figures 2f, 2g, 2h and 2i. The dry-stacked (mortarless) concrete blocks are
discrete units that transmit shear through concrete keys, interface friction, mechanical
connectors, or a combination of these methods. The stacked facing units result in potential failure planes through the facing column and this requires that additional stability
calculations be carried out to estimate interface shear forces and to compare these
forces with available shear capacity. In addition, the connection between the reinforcement layers and the facia is typically formed by extending the reinforcing layers along
the interface between facing units to the front of the wall. The connection detail must
also be evaluated for satisfactory design capacity (Bathurst and Simac 1993).
The NCMA method proposes a consistent approach to calculate earth pressures for
both external and internal stability calculations that is based on Coulomb earth pressure
theory. The advantage of Coulomb earth pressure theory over Rankine theory, which
has been adopted in earlier guidelines for geosynthetic-reinforced soil walls (AASHTO
1990, 1992; Christopher et al. (FHWA) 1989), is that the former explicitly accounts for
the influence of wall batter and wall-soil friction on the development of earth pressures
and hence is less conservative. A review of the essential features of the Coulomb earth
pressure approach as it applies to segmental retaining wall structures and comparisons
of the NCMA design methodology with earlier limit-equilibrium methods of design can
be found in the paper by Bathurst et al. (1993).
The scope of the NCMA guidelines is currently restricted to design of routine segmental retaining walls under static loading conditions. Questions related to the performance of the discrete facia system and the connections between the facia units and geosynthetic reinforcement layers during a seismic event have been raised (Allen 1993).
Nevertheless, the satisfactory performance of a number of geosynthetic-reinforced segmental walls during the Loma Prieta Earthquake of 1989 (Eliahu and Watt 1991; Collin
et al. 1992) and the Northridge Earthquake of 1994 in California (Sandri 1994) has been
qualitatively demonstrated.
The present paper investigates the stability of geosynthetic-reinforced segmental retaining walls under dynamic loading due to seismic excitation (earthquake). The study
is restricted to structures built on competent foundations for which foundation collapse
or excessive settlement is not a potential source of instability. A pseudo-static rigid
body approach that uses the well-known Mononobe-Okabe (M-O) method to calculate
dynamic earth forces (Okabe 1924) is outlined in this study. The method is restricted
to a limit-equilibrium approach in which factors of safety against collapse or rupture
mechanisms are calculated. The method developed in the current study should not be
confused with displacement methods which have been used for the design of conventional walls and explicitly incorporate permanent displacement criteria in stability
analyses (Richards and Elms 1979; Whitman 1990). The current approach is a logical
extension of the Coulomb wedge theory adopted by the NCMA guidelines for structures
under static loading. The M-O method has been used to calculate earth forces for seismic stability analyses of conventional gravity wall structures (Seed and Whitman 1970;
Richards and Elms 1979). Some of the questions raised in earlier papers concerned with
the implementation of pseudo-static methods for conventional gravity retaining walls
must also be addressed for the special class of structure that is the focus of this paper.

790

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

MONONOBE-OKABE EARTH PRESSURE THEORY

2.1

Calculation of Dynamic Earth Force

The Mononobe-Okabe method is used to calculate dynamic active earth forces acting
on a planar surface that is inclined at an angle, , into an unsaturated, homogeneous,
cohesionless soil mass (Figure 3). In Figure 3, W refers to the static weight of the active
wedge of soil acting behind the wall and Ww refers to the static weight of the facing column. Quantities kh and kv are horizontal and vertical seismic coefficients, respectively,
expressed as fractions of the gravitational constant, g . In the current study, horizontal
inertial forces are assumed to act outwards (+kh ) to be consistent with the notion of active earth pressure conditions. The convention adopted in this paper is that a positive
vertical seismic coefficient, +kv , corresponds to a seismic inertial force that acts downward and a negative seismic coefficient, - kv , corresponds to a seismic inertial force that
acts upward. The total dynamic active earth force, PAE , imparted by the backfill soil
is calculated as (Seed and Whitman 1970):
P AE = 1 (1 kv)K AEH 2
2

(1)

where: = unit weight of the soil; and H = height of the wall. The dynamic earth pressure
coefficient, KAE , can be calculated as follows:

+kh W

PAE
H

kh Ww

(1 kv )W

(1 kv )Ww
AE

Figure 3. Forces and geometry used in pseudo-static seismic analysis.

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

791

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

K AE =

cos 2( + )

sin(+)sin()
cos(+)cos(+)

cos cos 2 cos( + ) 1 +

(2)

where: = peak soil friction angle; = total wall inclination (positive in a clockwise
direction from the vertical); = mobilized interface friction angle assumed to act at the
back of the wall; = backslope angle (from horizontal); and = seismic inertia angle
given by:
= tan 1

1 k k
h

(3)

The seismic inertia angle represents the angle through which the vectorial resultant
of the gravity force and the inertial forces (both horizontal and vertical) is rotated from
vertical. Equations 1 to 3 are an exact analytical solution to the classical Coulomb
wedge problem that is modified to include the inertial forces kh W and kv W. Examination
of the trigonometric terms in Equation 2 shows that solutions are only possible for
. Hence, the maximum value of horizontal seismic coefficient for which there
are solutions to Equation 2 is restricted to kh (1kv )tan(-- ) .
Equations 1 and 2 can be modified to account for additional loads due to a uniformly
distributed surcharge acting behind the wall (Okabe 1924; Motta 1994). However, the
influence of any surface distributed surcharge loading on the stability of segmental retaining walls is not investigated in the current study. A closed-form solution for the calculation of dynamic earth force for c-- soils in retaining wall design is reported by Prakash (1981); however, this solution is restricted to the special case of = 0 and kv = 0.
For more complicated wall geometries and cases with surface loadings, trial single failure plane geometries, or two-part wedge failure plane geometries, can be evaluated to
find the critical geometry giving the maximum value of PAE . However, while these solutions are more general, they do not offer the designer the convenience of the closedform solutions adopted in the current study.
In the discussions to follow, it is convenient to decompose the total dynamic active
earth force, PAE , calculated according to Equations 1 and 2 into two components representing the static earth force component, PA , and the incremental dynamic earth force
due to inertial seismic effects, Pdyn (Seed and Whitman 1970). Hence:
P AE = P A + P dyn

(4)

(1 k v)KAE = K A + K dyn

(5)

or

where: KA = static active earth pressure coefficient; and Kdyn = incremental dynamic
active earth pressure coefficient. For brevity in the following text, the quantity PAE will
be called the dynamic earth force.

792

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

2.2

Distribution of Dynamic Lateral Earth Pressures and Point of Application

The position of the dynamic earth force, PAE , acting against gravity retaining walls
has been shown to be variable and to depend on the magnitude of ground acceleration.
A general range for the point of application of the dynamic force increment (Pdyn in
Equation 4) has been reported to be H=0.4H to 0.7H above the toe of the wall (Seed
and Whitman 1970). ( is the distance of the dynamic load increment above the toe of
the wall normalized with respect to wall height, H.) Seed and Whitman suggest that a
value of = 0.6 is reasonable for practical design purposes and this value is consistent
with the results of small-scale shake table tests reported by Ishibashi and Fang (1987).
Based on experience with conventional gravity retaining walls the earth pressure distributions illustrated in Figure 4 for = 0.6 are assumed to be applicable to segmental
retaining wall structures in this paper. The parameter m in Figure 4 denotes the normal-

0.8Kdyn H

0.8Kdyn H

Pdyn

PA

PAE = PA + Pdyn

H
mH

H/3

KA H
(a) static component

0.2Kdyn H
(b) dynamic increment

(KA +0.2Kdyn )H
(c) dynamic (total) pressure
distribution

Figure 4. Calculation of dynamic earth pressure distribution due to soil self-weight.


(Note: =0.6 .)

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

793

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

ized point of application of the dynamic earth force and is limited to the range
1/3m0.6. This range compares favorably to measured values of m ranging from 0.3
to 0.5 reported by Ichihara and Matsuzawa (1973) from the results of shake table tests
on small-scale gravity wall models. The distributions for static and dynamic increment
of active earth pressure illustrated in Figure 4 are also identical to those recommended
for the design of flexible anchored sheet pile walls (Ebling and Morrison 1993).
Finally, to simplify all stability calculations in this paper, and to be consistent with
the convention adopted in the NCMA guidelines, only the horizontal component of PAE
is used in stability calculations, i.e. PAE cos(-- ). This assumption results in a conservative (i.e. safe) design by ignoring the stabilizing benefit of the vertical component of
PAE .
2.3

Orientation of Active Failure Plane

Closed-form solutions for the orientation of the critical planar surface from the horizontal, AE , have been reported by Okabe (1924) and Zarrabi (1979). These solutions
are rewritten here as:
a AE = + tan1

A E + D
AE

AE

AE

(6)

where
A AE = tan( )
D AE = A AEA AE + B AE B AECAE + 1
E AE = 1 + C AE A AE + B AE
B AE = 1tan( + )
C AE = tan( + )
Equation 6 can be used to calculate the orientation of the assumed active failure plane
within the reinforced soil mass and in the retained soil.
2.4

Selection of Parameter Values

2.4.1 Soil and Interface Friction Angles


In the theoretical developments and parametric analyses to follow, the friction angle,
, of the cohesionless backfill soils is assumed to be the peak value determined from
conventional laboratory practice and its magnitude is assumed to be unchanged under
seismic excitation due to an earthquake. The choice of peak friction angle for seismic

794

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

design is consistent with AASHTO (1990, 1992), FHWA (Christopher et al. 1989) and
NCMA (Simac et al. 1993) guidelines for static design of geosynthetic-reinforced soil
walls. Under rapid loading conditions the strength of compacted unsaturated cohesionless backfills can be expected to be at least as great as the static value.
The admissible range of the interface friction angle, , is 0 in Coulomb wedge
analyses. In static stability analyses, is assumed to be equal to 2/3 for internal stability analyses (facing column-reinforced soil interface) and = for external stability
analyses (reinforced soil-retained soil interface). A value of 2/3 has been shown to be
applicable for wall-soil interface friction based on small-scale shake table tests of conventional gravity wall structures (Ishibashi and Fang 1987) and is assumed to be also
applicable for geosynthetic-reinforced retaining wall structures in the current study.
Parametric analyses are restricted to the case to avoid the complication that results from vertical components of earth forces that act upward. The condition is
valid for typical segmental retaining walls since fully-mobilized interface friction
angles, , taken at the facing column-reinforced soil interface and the reinforced soilretained soil interface are generally greater than the wall inclination angle, .
2.4.2 Seismic Coefficients
The general solution to the M-O method of analysis admits both vertical and horizontal components of seismic-induced inertial forces. The range of horizontal seismic coefficients used in the parametric analyses to follow is restricted to kh < 0.5 .
The choice of positive or negative kv values will influence the magnitude of dynamic
earth forces calculated using Equations 1 and 2. In addition, the resistance terms in factor of safety expressions introduced later in the paper will be influenced by the choice
of sign for kv . An implicit assumption in many of the papers on pseudo-static design
of conventional gravity wall structures reviewed by the writers is that the vertical component of seismic body forces acts upward. However, the designer must evaluate both
positive and negative values of kv to ensure that the most critical condition is considered
in dynamic stability analyses if non-zero values of kv are assumed to apply. For example,
Fang and Chen (1995) have demonstrated in a series of example calculations that the
magnitude of PAE may be 12% higher for the case when the vertical seismic force acts
downward (+kv ) compared to the case when it acts upward (-- kv ). Nevertheless, selection
of a non-zero value of kv implies that peak horizontal and vertical accelerations are time
coincident which is an unlikely occurrence in practice. The assumption that peak vertical accelerations do not occur simultaneously with peak horizontal accelerations is
made in the current FHWA guidelines for the seismic design of mechanically stabilized
soil retaining walls (Christopher et al. 1989). Indeed, Seed and Whitman (1970) have
suggested that kv = 0 is a reasonable assumption for practical design of conventional
gravity structures using pseudo-static methods. Wolfe et al. (1978) studied the effect
of combined horizontal and vertical ground acceleration on the seismic stability of reduced-scale model Reinforced Earth walls using shake table tests. They concluded that
the vertical component of the seismic motion may be disregarded in terms of practical
seismic stability design. Their conclusion can also be argued to apply to geosyntheticreinforced segmental walls. Nevertheless, significant vertical accelerations may occur
at sites located at short epicentral distances and engineering judgement must be exer-

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

795

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

cised in the selection of vertical and horizontal seismic coefficients to be used in pseudo-static seismic analyses.
However, in order to address specific concerns raised by Allen (1993) related to facing stability of geosynthetic-reinforced segmental retaining walls during a seismic
event that includes vertical ground accelerations, parametric analyses were carried out
in the current study for the range kv = - 2kh /3 to + 2kh /3. The upper limit on the ratio kv
to kh is equal to the calculated ratio of peak vertical ground acceleration to peak horizontal ground acceleration from seismic data recorded in the Los Angeles area
(UCB/EERC, 1994).
In conventional pseudo-static M-O methods of analysis the choice of horizontal seismic coefficient, kh , for design is related to a specified horizontal peak ground acceleration for the site, ah . On the west coast of British Columbia (the most seismically active
area of Canada), the typical maximum design horizontal ground acceleration on rock,
based on a 10% probability of exceedance in 50 years, is ah = 0.32g (CFEM 1992). The
relationship between peak ground acceleration for the site, ah , and a representative value of kh is nevertheless complex and there does not appear to be a general consensus
in the literature on how to relate these parameters. For example, Whitman (1990) reports that values of kh from 0.05 to 0.15 are typical values for the design of conventional
gravity wall structures and these values correspond to 1/3 to 1/2 of the peak acceleration
of the design earthquake. Bonaparte et al. (1986) used kh = 0.85ah /g to generate design
charts for geosynthetic-reinforced slopes under seismic loading using the M-O method
of analysis. However, the results of finite element (FE) modelling of reinforced soil
walls by Segrestin and Bastick (1988), Cai and Bathurst (1995) and limited 1/2 scale
experimental work (Chida et al. 1982) has shown that the average acceleration of the
composite soil mass may be equal to or greater than ah depending on a number of factors
such as: magnitude of peak ground acceleration; predominant modal frequency of
ground motion; duration of motion; height of wall; and stiffness of the composite mass.
Current FHWA guidelines use an equation proposed by Segrestin and Bastick (1988)
that relates kh to ah according to:
k h = (1.45 a hg)(ahg)

(7)

This formula results in kh > ah /g for ah < 0.45g . However, as clearly stated by Segrestin
and Bastick, their equation should be used with caution because it is based on the results
of FE modeling of steel reinforced soil walls up to 10.5 m high that were subjected to
ground motions with a very high predominant frequency of 8 Hz. The results of FE modeling reported by Cai and Bathurst (1995) for a 3.2 m high geosynthetic-reinforced segmental retaining wall with ah = 0.25g and a predominant frequency range of 0.5 to 2
Hz gave a distribution of peak horizontal acceleration through the height of the composite mass and retained soil that was for practical purposes uniform and equal to the base
peak input acceleration. These observations are consistent with the results of Chida et
al. (1982) who constructed 4.4 m high steel reinforced soil wall models and showed that
the average peak horizontal acceleration in the soil behind the walls was equal to the
peak ground acceleration for ground motion frequencies less than 3 Hz. In practice, the
final choice of kh may be based on local experience, and/or prescribed by local building
codes or other regulations.

796

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

In the current analyses, kh and kv are assumed to be uniform and constant throughout
the facing column, the reinforced soil mass and in the retained soils. This assumption
simplifies the analysis for geosynthetic-reinforced soil walls but may not be true for
walls higher than (say) 7 m, or walls with complex geometries, surface loadings, and/or
difficult foundation conditions. For these structures, and/or structures subjected to high
frequency ground motions, more sophisticated analyses may be warranted.
2.4.3 Other
In order to simplify stability analyses in this study, the facing units and backfill soils
are assumed to have the same unit weight, . This assumption introduces negligible
error in the example calculations to follow. It can be noted that the majority of the segmental retaining wall units on the market today are hollow soil infilled masonry units.
For these systems the facing unit weights and typical backfill soil unit weights are very
similar.
2.5

Parametric Analyses Related to Mononobe-Okabe Earth Pressure Theory

A distinguishing feature of the geometry of segmental retaining walls is that the surfaces against which active earth pressures are assumed to act are oriented at > 0 from
the vertical (Figure 3). Hence, the wall-soil interface and reinforced soil-retained soil
interface are rotated in the opposite direction to that of many conventional gravity wall
structures. Typically, varies from 3 to 15_ from the vertical (Simac et al. 1993) depending on the setback of the stacked modular units and the initial base unit inclination.
The influence of horizontal seismic coefficient, kh , and wall inclination angle, , on
the dynamic earth force, PAE , is illustrated in Figure 5. The data shows that the effect
of positive wall inclination is to reduce dynamic earth forces to levels less than those
developed against conventional gravity wall structures of the same height and retaining
the same frictional soil.
The backslope angle, , also influences the magnitude of dynamic earth force. As the
backslope angle becomes larger, the magnitude of dynamic earth force increases as illustrated in Figure 6. The effect of increasing positive wall inclination is seen to reduce
the magnitude of dynamic earth force for a given backslope angle.
The influence of wall inclination angle on the orientation of the critical failure surface, AE , through the reinforced soil mass is shown in Figure 7. The figure illustrates
that for a given wall inclination angle, the size of the active soil wedge behind the wall
facing increases with increasing magnitude of the horizontal seismic coefficient, kh . A
similar result has been reported by Vrymoed (1989). The implication of the results
shown in Figure 7 to internal stability design is that the length of reinforcement layers,
particularly those near the top of the reinforced soil zone, may have to be extended in
order to capture potential failure surfaces propagating from the heel of the facing column (inset diagram in Figure 7). A similar conclusion was made by Bonaparte et al.
(1986) for the design of reinforced slopes under seismic loading. The same calculation
for AE can be carried out to determine the critical failure plane through the retained soil
mass that is assumed to propagate from the heel of the reinforced soil mass. The implication of such calculations to external stability analysis is that the size of the failure

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

797

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

1.2

---15_

1.1

0.9

PAE

0.8
2P AE
H 2

---10_

1.0

kv

= 35_
= 2/3
= 0_
=0

--- 5_
0_
+ 5_

0.7
0.6

+10_
+15_
Conventional walls

0.5
0.4
0.3
0.2
0.1
0.0
0.0

Segmental retaining walls


0.1

0.2

0.3

0.4

0.5

kh
Figure 5. Influence of seismic coefficient, kh , and wall inclination angle, , on dynamic
earth force, PAE .

zone behind the reinforced soil zone will also increase with increasing horizontal acceleration.
The combined influence of horizontal and vertical accelerations on dynamic earth
force, PAE , is illustrated in Figure 8 for two wall inclination values. For values of horizontal seismic coefficient, kh , less than about 0.35 and kv = 2kh /3, downward vertical
components of seismic earth force (kv > 0) give the largest dynamic earth forces. However, the value of PAE using kv = +2kh /3 is only 7% greater than the value calculated using
kv = 0 for kh < 0.35. Hence, over a wide range of horizontal seismic coefficient values
the assumption that kv = 0 is reasonably accurate and, in fact, results in a slightly more
conservative value of PAE than values calculated assuming that the vertical component
of seismic earth force acts upward (kv < 0).

798

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

1.4

= 10_

1.3

1.1

0_

1.2
PAE

1.0

5_
= 20_

0.9

0_

0.8

5_

10_
15_

= 0_

0_
5_

2P AE 0.7
H 2

10_

0.6

15_

10_
15_

0.5
0.4
= 35_
= 2/3
kv = 0

0.3
0.2
0.1
0.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

kh
Figure 6. Influence of seismic coefficient, kh , backslope angle, , and wall inclination angle,
, on dynamic earth force, PAE .

The influence of backslope angle, , and wall inclination angle, , on calculated values of dynamic earth force can be argued to have a more significant effect than the magnitude of kv in the example calculations. The requirement that peak horizontal and vertical accelerations must be time coincident in order to generate even the small calculated
differences in dynamic earth forces shown here gives support to recommendations in
current design guidelines for conventional gravity structures that vertical accelerations
can be safely ignored in many seismically active areas.

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

799

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

70

kv

---15_ 65
---10_
--- 5_
60
0_
+ 5_

=
=
=
=

35_
2/3
0_
0

Conventional walls

AE (degrees)

+10_ 55
+15_
50
45
40

Segmental
retaining walls
+

35

kh = 0

kh > 0

30
25
20
0.0

Increased
reinforced soil mass
for kh > 0
Increased reinforcement
lengths

AE

Reinforced soil mass for kh = 0


0.1

0.2

0.3

0.4

0.5

kh
Figure 7. Influence of seismic coefficient, kh , and wall inclination angle, , on orientation
of internal failure plane, AE .

EXTERNAL STABILITY ANALYSES

3.1

General

Potential external modes of failure are illustrated in Figures 2a, 2b and 2c. The analyses described in this section assume that the foundation provides a competent base such
that potential modes of failure are restricted to translational sliding along the base and

800

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

1.0

0.9

PAE

0.8
H
0.7
0.6
2P AE 0.5
H 2

= 0_
kv = +2kh /3
kv = 0
kv = ---2kh /3

0.4
0.3

= 10_
kv = +2kh /3

0.2

kv = 0
0.1
0.0
0.0

= 35_
= 2/3
= 0_
0.1

kv = ---2kh /3

0.2

0.3

0.4

0.5

0.6

kh
Figure 8. Influence of seismic coefficients, kh and kv , and wall inclination angle, , on
dynamic earth force, PAE .

overturning about the toe of the gravity mass (reinforced soil zone plus facing column).
The dynamic earth force, PAE , calculated according to Equation 1 is used in external
stability calculations to estimate destabilizing active earth forces. In the external stability analyses to follow, the dynamic earth force imparted by the cohesionless retained
soil on the gravity mass is assumed to act along a surface that is parallel to the wall face
(i.e. at angle from the vertical) and at a constant distance L from the front face of the
wall (i.e. L is the minimum width of the gravity mass). According to NCMA guidelines,

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

801

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

the minimum width of the gravity mass is L = 0.6H for critical structures and L = 0.5H
for non-critical structures. Dimension L is used to locate the heel of the assumed minimum gravity mass for external stability calculations and to control the minimum length
of all reinforcement layers (Figure 9).
3.2

Base Sliding (Figure 2a)

The dynamic factor of safety against base sliding for purely frictional soils can be expressed as:

12
11

FS bsl (static)

kv = ---2kh /3

10

kv = ---kh /2

kh WR
WR (1kv )
Lw

PAE cos()

kv =0

6
5
4
3

= 20_
kv = ---2kh /3
kv = ---kh /2
kv = ---kh /4
kv = 0

= 0_

2
1.5

kv = ---kh /4

R = WR (1kv )tan

1
0
0.0

0.1
0.06

0.2

0.3

=
=
=
=
=

0.6
35_

0_
0.5H

0.4

0.5

0.6

kh

Figure 9. Static factor of safety against base sliding to give a minimum dynamic factor of
safety of 1.125 against base sliding for a range of seismic coefficients, kh and kv , and
backslope angle, .
(Notes: WR = weight of reinforced zone plus weight of facing column; and R = base sliding resistance.)

802

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

L
LL
a + (1 k ) tan
H
H
LL
L
(1 k )a cos( ) + k
a +
H
H
w

FS bsl =

1K
2 AE

2
1

(8)

where:
a1 = 1 +

L Lw
tan
H

a2 = 1 +

L Lw
tan
2H

Here, Lw is the width of the facing column. Equation 8 can be simplified by setting Lw
= 0 for Lw << L. This assumption introduces no error for = 0 and negligible error for
moderate values of > 0. Parameter is an empirical constant that is used to artificially
reduce the inertial force of the gravity mass and applies only to the inertial part of Equation 8. A value of = 0.6 has been used for design purposes for both geosynthetic-reinforced soil walls (Christopher et al. 1989) and for Reinforced Earth walls that use steel
reinforcement strips (Segrestin and Bastick 1988). Parameter is assumed to be less
than unity to account for the transient nature of the peak accelerations in the gravity
mass and retained soils and the expectation that the inertial forces induced in the gravity
mass and the retained soil zone will not reach peak values at the same time during a
seismic event. The terms a1 and a2 in Equation 8 are geometric constants that account
for the effect of the backslope angle on the calculation of the mass of the reinforced soil
zone. The factor of safety for static loading conditions (FSbsl (static)) can be recovered
from Equation 8 by setting kh = kv = 0 .
The FHWA (Christopher et al. 1989) recommends that the minimum factor of safety
against base sliding under dynamic loading be no less than 75% of the minimum allowable static value. If this rule is applied to a minimum allowable static factor of safety
against sliding of 1.5 then, the data in Figure 9 shows that the maximum permissible
ground acceleration is 0.06g. Larger static factors of safety will be required to ensure
FSbsl (dynamic) (0.75)(1.5) = 1.125 as illustrated in the figure. Nevertheless, there
are limiting seismic coefficient values for any set of wall parameters for which there
is no solution using Equations 1 and 2 (i.e. < 0 ). The analytical results described here are consistent with conclusions made by Richards and Elms (1979) who
showed that there is very little margin of safety against base sliding of gravity structures
designed to meet minimum conventional static factors of safety. Figure 9 also demonstrates that the required static factor of safety against base sliding for a structure with
a horizontal backslope is relatively insensitive to the magnitude of vertical acceleration
for a horizontal seismic coefficient kh < 0.3. The vertical seismic coefficient used to generate the curves in Figure 9 was taken as kv 0 since negative values of kv gave the
most conservative results (i.e. largest required FSsld (static) values). The data in Figure
9 do show that even a modest backslope angle will require large increases in the static
factor of safety against base sliding to ensure an acceptable margin of safety under seismic loading.

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

803

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

3.3

Overturning (Figure 2b)

The moment arm, Ydyn , of the dynamic force normalized with respect to the wall
height (Figure 4) can be calculated as follows:
1 K + [K (1 k ) K ]
v
A
AE
A
Y dyn
m=
=3
H
KAE(1 k v)

(9)

The relationship between normalized moment arm, m, and horizontal seismic coefficient, kh , is shown in Figure 10a for values of ranging from 0.4 to 0.7 and = and
= 2/3 . The location of the dynamic force moment arm, m, is sensitive to the assumed
value of but is relatively insensitive to the range of interface friction angle, . For
0.6, the point of application of the dynamic force is generally at 1/3 to 1/2 of the wall
= 35_
0.70
0.65

=
= 2/3

0.7

0.60
0.55
m

= 0_
0.70

0.6

0.50

kv = 0

0.65
0.60

0.55

15_
0_
---15_

0.50
0.5

0.45
0.40

0.4

0.35

0.45
0.40
0.35

0.30
0.30
0.0 0.1 0.2 0.3 0.4 0.5 0.6
0.0 0.1 0.2 0.3 0.4 0.5 0.6
kh

kh
(a) = 0_

(b) = 0.6, = 2/3

Figure 10. Influence of seismic coefficient, kh , normalized dynamic force increment


location, , wall inclination angle, , and wall-soil interface friction angle, , on location of
normalized dynamic moment arm, m.

804

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

height above the toe. Figure 10b illustrates the influence of horizontal seismic coefficient, kh , and magnitude of wall inclination angle, , on parameter m for = 0.6. The
point of application of the dynamic earth force is shown to be only slightly dependent
on the magnitude of wall inclination angle associated with typical segmental retaining
walls.
The dynamic factor of safety against overturning about the toe of the free body comprising the reinforced soil mass and the facing column can be calculated as follows:

LL w
H

FS bot =

L
2L
b + LL
a + (1 k )
H
2

LL w
L
mKAE(1 k v)a 31 cos( ) + k h
b1 + w
H
H

(10)

where
b 1 = a1 + 1 (a1 1)2
3
b 2 = 1 + 2 (a 1 1)
3
and terms a1 and a2 are defined in Equation 8. Equation 10 is applicable to the case of
a vertical wall ( = 0) and will give a slightly conservative estimate of FSbot for inclined
wall facing columns. The static factor of safety, FSbot (static), required to satisfy FSbot
(dynamic) = (0.75)(2.0) = 1.5 is plotted in Figure 11 for the case of two different reinforcement length to wall height ratios, L/H . The vertical component of seismic force
has been taken as upward (-- kv ) in order to calculate results for the most critical orientation. If the conventional rule that the dynamic factor of safety be not less than 75% of
the minimum allowable static factor of safety against overturning (i.e. 2) is applied,
then the horizontal seismic coefficient is limited to kh = 0.04 in the example calculations. The data in Figure 11 illustrates that even a modest backslope angle leads to a
dramatic increase in the required static factor of safety for a given design acceleration.
Indeed, for the case of = 20_ and kh > 0.25 an acceptable margin of safety against overturning may not be possible. Figure 11 also shows that the effect of the L/H ratio on the
required value of FSbot (static) is relatively insignificant compared to the effect of backslope inclination for typical structures designed for static loading environments.
4

INTERNAL STABILITY

4.1

General

The influence of the magnitude of seismic coefficients on lateral earth forces has been
demonstrated earlier in the paper. If the calculation of dynamic reinforcement loads is
carried out in the same manner as for conventional static wall design then, the effect
of seismic loading can be shown to increase the magnitude of the net horizontal force
carried by the geosynthetic reinforcement layers. In addition, the change in the distribu-

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

805

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

kv = ---2kh /3

15

13
12

Lw

14

kv = 0

kh WR
H

11

ma2 H

WR (1kv )

10
FSbot (static)

= 0_

PAE cos()

kv = ---2kh /3

kv = 0

= 20_

L/H = 0.5

L/H = 0.7

6
5
4
3

2
1
0
0.0

0.1
0.04

0.2

0.3

=
=
=
=
=
0.4

0.6
35_

0_
0.6
0.5

kh

Figure 11. Minimum static factor of safety against overturning required to give a factor of
safety of 1.5 against dynamic overturning for a range of seismic coefficients, kh and kv ,
backslope angle, , and length to height ratio, L/H.
(Note: WR = weight of reinforced zone plus weight of facing column.)

tion of the lateral earth pressure (Figure 4) means that the percentage of total lateral
force to be carried by the reinforcing elements in the upper portions of the wall will increase. Finally, the influence of ground accelerations on the volume of the internal potential failure wedge leads to an increase in the length of the reinforcement layers as
discussed in Section 2.5 and illustrated in Figure 7.

806

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

4.2

Over-stressing of Reinforcement (Figure 2d)

The contributory area approach used for the static stability analysis of segmental retaining walls is extended to the dynamic loading case. In this method the reinforcement
layers are modelled as tie-backs with the dynamic tensile force, Fdyn , in each layer equal
to the dynamic earth pressure integrated over the contributory area, Sv , at the back of
the facing column plus the corresponding wall inertial force increment, kh Ww . The
contributory area for the topmost reinforcement layer is taken from the top of the crest
to mid-elevation between the first and second reinforcement layers from the crest. For
the simple geometry illustrated in Figure 12, the dynamic factor of safety, FSos , against
over-stressing of a reinforcement layer at depth z below the crest of the wall is given
by:
FS os =
=

T allow
F dyn

(11)

T allow

L
0.8K dyn cos( ) + (K A 0.6K dyn) cos( ) z + k h W HS v
H
H

Here, Tallow is the allowable tensile load for the reinforcement under seismic loading.

Dynamic
earth pressure
distribution

0.8Kdyn Hcos()

H
kh Ww

LW

Sv

Fdyn

(KA +0.2Kdyn )Hcos()

Reinforcement
layer (typical)

Figure 12. Calculation of tensile load, Fdyn , in a reinforcement layer due to dynamic earth
pressure and wall inertia.
(Note: = 0.6.)

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

807

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

The influence of seismic coefficient values on the magnitude of tensile reinforcement


loads can be explored by computing a magnification factor, rF , that is the ratio of dynamic tensile force, Fdyn , to static tensile force, Fsta , for a reinforcement layer at depth
z below the wall crest. Results of this calculation for reinforcement layers at five different depths below the wall crest are presented in Figure 13. The data illustrates that the
largest increases in reinforcement force occur in the shallowest layers in a reinforced
soil wall. This result is not unexpected due to the change in the active earth pressure
distribution that results from dynamic loading as discussed in Section 2.2 and illustrated
10
9
8
7

= 35_

0.2

F dyn
rF =
F sta

= 0_
= 0.6
= 2/3
= 0_

0.3

LW /H= 0.1
kv = ---2kh /3

kv = 0
kv = +2kh /3

rF

0.4

5
4
0.6
3
0.8
2
z/H
1
0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

kh
Figure 13. Influence of seismic coefficients, kh and kv , and normalized depth below crest
of wall, z/H, on dynamic reinforcement force amplification factor, rF .

808

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

in Figures 4 and 12. The plotted data in Figure 13 also reveals that rF is sensibly independent of the magnitude of kv for kh 0.35, and hence solutions using kv = 0 are sufficiently accurate for design over this range and even slightly conservative at the shallowest
depth investigated (z/H=0.2).
An implication of these results to design is that the number of reinforcement layers
may have to be increased at the top of the wall in order to keep tensile loads within allowable limits. However, it should be noted that the allowable design tensile load under
dynamic loading is routinely taken as a greater percentage of the index strength of the
reinforcement than the percentage used for static loading design because of the short
duration of peak tensile loading during a seismic event. AASHTO (1992) guidelines can
be interpreted to permit the value of Tallow used for static loading designs to be increased
by 33% for the seismic loading condition. Rapid in-isolation wide-width strip tensile
loading of a typical high density polyethylene (HDPE) geogrid reinforcement reported
by Bathurst and Cai (1994) has demonstrated a potentially large increase in reinforcement stiffness of the material when compared to conventional rates of loading. This observation suggests that HDPE geogrids may be designed for much greater strengths under seismic loading than those values that result from the interpretation of AASHTO
recommendations.
4.3

Reinforcement Anchorage (Figure 2e)

The dynamic reinforcement tensile load must be carried by the reinforcement anchorage length which is located between the internal active failure plane (oriented at AE
from horizontal) and the reinforcement free end (Figure 7). A common approach for
anchorage capacity design is to use a simple Coulomb-type interface model in which
anchorage capacity is linearly proportional to anchorage length, overburden pressure
and soil shear strength (AASHTO 1990). An implication of this model for dynamic anchorage design is that the required anchorage length will increase in proportion to the
magnification factor, rF , introduced in the previous section. Owing to the short duration
of the anchorage force during a seismic event, the factor of safety against anchorage
pullout may be taken as 75% of the static value of 1.5 according to AASHTO (1992).
However, the principal effect of dynamic loading on anchorage design is the requirement to increase the length of reinforcement layers in order to capture the larger active
wedge of soil that occurs as a result of seismic loading, i.e. the internal failure plane
angle, AE , decreases as kh increases. This effect is illustrated in Figure 14. For example,
this figure shows that for a horizontal seismic coefficient, kh = 0.25, reinforcement
lengths would have to be increased by 60 to 100% of the lengths required under static
loading conditions. The requirement that the lengths of the uppermost reinforcement
layers may need to be increased for reinforced slopes subject to seismic loading has
been noted by Bonaparte et al. (1986) and is based on similar arguments.
4.4

Internal Sliding (Figure 2f)

The modular facing construction in geosynthetic-reinforced segmental retaining


walls requires that the designer check for internal sliding along horizontal planes that
pass along the reinforcement-soil interface and through the facing column between facing units. The results of large-scale shear tests reported by Bathurst and Simac (1994)

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

809

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

5.0

= 30_ 35_ 40_ 45_


30_

kv = 0
kv = ---2kh /3

4.5
+
4.0

kh = 0

kh > 0

L sta

3.5

L dyn

35_

AE

L dyn 3.0
Lsta

40_
45_

2.5

2.0
1.6

= 0_
1.5

= 0_
= 2/3

0.25
1.0
0.0

0.1

0.2

0.3

0.4

0.5

kh
Figure 14. Influence of seismic coefficients, kh and kv , and soil friction angle, , on ratio of
minimum reinforcement lengths, Ldyn /Lsta , to capture the internal failure wedge in
pseudo-static Coulomb wedge analyses.

have shown that the static shearing resistance, Vu , available at a horizontal interface
in the facing column can be described by a Coulomb-type failure law. This failure criterion can be modified to account for the dynamic loading case as follows:

810

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

V u = au + Ww(1 k v) tan u

(12)

Parameters, au and u represent minimum interface shear capacity and equivalent interface friction angle, respectively, and are not expected to vary between static and seismic
loading conditions. The analysis required to calculate the factor of safety against internal sliding is similar to that described for external sliding. The dynamic factor of safety
against internal sliding along a horizontal surface at depth z below the crest of the wall
is calculated as:

FS isl =

Vu
LLw
+
a 2 (1 k v) tan ds
z
z2

LLz a + Lz

1 K (1 k )a 2 cos( ) + k
v 1
h
2 AE

(13)

Coefficients a1 and a2 are the same as those reported for Equation 8 substituting H = z.
Equation 13 assumes that the critical internal sliding mass is initiated at the free end
of the reinforcement layer. Parameter ds is the interface (direct sliding) friction angle
between the geosynthetic-reinforcement and the cohesionless reinforced soil. In general, ds < but the reduction in sliding resistance is typically more than compensated
for by the large value of shear interlock that is available in many block systems. However, the combined effect of a low interface friction value, ds , and facing units with low
shear capacity, Vu , can result in unacceptably low factors of safety against interface
shear and this failure mechanism must be checked as a matter of routine.
A parametric analysis involving Equation 13 was not performed. However, the results
plotted in Figure 9 are applicable for internal sliding for the case au = 0, u = ds = and
setting H = z.
5

FACING STABILITY

5.1

Facing Failure Mechanisms

The following potential failure mechanisms must be examined in pseudo-static seismic analysis of the facing stability of segmental retaining walls: interface shear failure;
connection failure; toppling (local overturning) (Figure 2).
5.2

Interface Shear (Figure 2g)

The influence of interface shear transmission on facing column stability can be analyzed by treating the facing column as a beam in which the integrated lateral pressure
(i.e. distributed load) must equal the sum of the reactions (forces in reinforcement layers). The calculation of interface shear force under dynamic loading must include the
effect of wall facing inertia. The general approach is illustrated in Figure 15. The total
force carried by reinforcement layers located above facing unit j is calculated as the area
ABDC of the lateral earth pressure distribution, plus the facing column inertial force

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

811

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

Horizontal
component of
dynamic
earth pressure
distribution

z
kh

W
j+1

j
w

i
F dyn

i+1

Svi+1
k h Wjw
S vi

j
S dyn

C
F

D
E

i
F dyn

j
S dyn
= k h Wjw + AREA CDEF

Figure 15. Calculation of dynamic interface shear force acting at a reinforcement


elevation.
(Notes: N = total number of reinforcement layers; and M = total number of facing units.)

over the same height. The out-of-balance force to be carried through shear at the bottom
of facing unit j is simply the sum of the incremental column inertial force k hWjw plus
the force due to area CDEF in the figure. The partitioning of forces illustrated in the
figure is a direct result of the contributory area approach introduced earlier to assign
tensile loads to reinforcement layers. The locally maximum interface shear forces will
occur at reinforcement elevations. A general expression for the factor of safety against
dynamic interface shear failure (FSsc (dynamic)) at a reinforcement layer is:
FS sc =
=

Vu
S dyn

0.8K

(14)
Vu
dyn cos(

S
) + (K A 0.6K dyn) cos( ) z v

+ k LH HS2

4H

where: Sdyn = interface shear force; and Vu = shear capacity. For facing columns
constructed at large inclination angles the magnitude of the normal force, Ww , transmitted between facing units may be less than the sum of the weights of the individual

812

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

facing units above the interface elevation. The loss of normal load is due to the effect
of the facing column units leaning into the reinforced soil mass. The maximum height
of column units that will transmit all of the facing weight to a target interface is called
the hinge height. Its calculation is described in detail by Bathurst et al. (1993) and
Simac et al. (1993).
The calculation of factor of safety against interface shear failure under static loading
conditions is carried out using Equation 14 with kv = kh = 0 . Figure 16 shows the ratio

1.0

= 35_

0.9

Sv /H = 0.2

= 0.6

0.8

FSsc (dynamic) / FS sc (static)

LW /H = 0.1

= 0_
= 2/3

kv = ---2kh /3

= 0_

kv = 0

0.7
0.6
z/H

0.5

0.9

0.4
0.7
0.3

0.9

0.2

0.7 0.5

0.1

0.5

0.0

0.1

0.3

0.3
0.0

0.1

0.2

0.3

0.4

0.1

0.5

0.6

kh
Figure 16. Influence of seismic coefficients, kh and kv , and normalized depth below crest
of wall, z/H, on the ratio of dynamic to static interface shear factor of safety.

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

813

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

of dynamic factor of safety to static factor of safety against interface shear failure for
a range of horizontal and vertical seismic coefficient values applied to an example wall
with five evenly spaced reinforcement layers (i.e. Sv /H=0.2). The data shows that the
potential for interface shear failure under seismic loading increases with proximity of
the shear interface to the crest of the wall. However, the effect of vertical acceleration
on calculated dynamic factors of safety diminishes with height of interface for the worst
case situation of vertical seismic force components acting upward. The curves for
z/H=0.1 correspond to sliding stability of the top unreinforced portion of the wall facing. These curves appear to support the argument that narrow unreinforced heights of
segmental facing units are susceptible to sliding failure. However, in practice, large values of interface shear capacity are possible with many modular block systems that are
constructed with shear keys, or other forms of positive interlock. These systems, as opposed to systems that rely solely on frictional sliding resistance, are the preferred choice
in order to achieve an adequate margin of safety against interface sliding. Furthermore,
Bathurst and Simac (1994) have demonstrated that the shear capacity at an interface
layer may be substantially reduced by the presence of a reinforcement inclusion if interface shear capacity is developed primarily through frictional resistance.
A minimum factor of safety against interface shear failure, for critical structures under static loading, is 1.5 according to NCMA guidelines. Reducing this criterion by 25%
for the dynamic loading condition is recommended in order to be consistent with the
approach adopted for base sliding.
5.3

Connections (Figure 2h)

The influence of increased dynamic forces on connection load is identical to the analysis described for reinforcement over-stressing. Peak connection load capacities under
static loading conditions have been described using bi-linear failure envelopes based
on the results of full-scale connection tests carried out at rates of loading matching the
10% strain/min rate used in the ASTM D 4595 method of test (Bathurst and Simac
1993). A Coulomb-type law with a maximum connection load cut-off has been used by
the first writer and co-workers to characterize a large number of (static) test results. Modified for the dynamic loading condition, the peak connection load envelope becomes:
F c = acs + Ww (1 k v) tan cs F c(max)

(15)

where the parameters acs and cs represent the minimum connection capacity and the
slope of the connection strength envelope, respectively. The dynamic factor of safety
against connection failure, FScn , is expressed by Equation 11 by replacing Tallow with
Fc . Depending on the connection type, it is possible that the maximum reinforcement
load, Fdyn , may be limited by the facing connection capacity. The results of connection
tests carried out at different rates of loading have demonstrated that peak connection
capacities may be sensitive to rate of loading (Bathurst and Simac 1993). At the time
of writing there is no data available that can be used to quantify changes in connection
capacity that may develop as a result of repeated application of load and the rapid loading rates anticipated during a seismic event.

814

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

5.4

Toppling (Figure 2i)

The distribution of internal moments at interface elevations can also be calculated


using the beam analog described in Section 5.2 for interface shear stability calculations.
Internal moments that cause net outward moment at the toe of a facing unit provide a
possible failure mechanism for which an adequate factor of safety should be checked.
Local peak destabilizing moments will occur at reinforcement elevations. The factor
of safety expression adopted by the NCMA (Simac et al. 1993), for local overturning
at a reinforcement layer i under static loading conditions, can be modified for the dynamic loading case as follows:
N

M R(1 k v) +
FS lot =

i
c

Y ic

i+1

1 K cos( ) z + (0.4 0.1 z )K cos( ) + 1 k L w Hz 2


dyn
H
H
2 hH
6 a

(16)

where: MR = resistance to static overturning due to facing column self-weight above the
toe of the target facing unit; and N = number of reinforcement layers. The summation
with F ic Yic terms denotes the resisting moment due to the connection capacities of reinforcement layers, F ic , and their corresponding moment arms, Y ic , from the target point
of rotation.
The ratio of the dynamic to static factor of safety against local overturning is plotted
in Figure 17 for five different interface elevations ranging from z/H = 0.2 to 1. To simplify example calculations the connection capacities, F ic , have been assumed to be purely
frictional (acs = 0). The largest reductions in factor of safety under seismic loading were
found to occur when the vertical component of the seismic force acts upward (kv =
- 2kh /3). The figure clearly shows that for the seismic coefficient values investigated,
the shallow interface layers (small z/H values) require a higher static factor of safety
against overturning to maintain a dynamic factor of safety equal to or greater than unity.
The unsupported height of the facing column at the top of the structure is the most critical portion of the wall in these calculations (i.e. z/H=0.2 in this example). However, the
effect of the magnitude and orientation of the vertical component of seismic force on
the upper portions of the wall (z/H < 0.4) in the example used to produce Figure 17 is
only significant for horizontal seismic coefficient values greater than 0.3.
The experience of the writers is that designers typically try to maximize the unreinforced height of wall at the crest in order to reduce reinforcement quantities in segmental retaining wall structures. This strategy will result in unacceptably low margins
of safety against toppling at the top of the structure under dynamic loading conditions
as illustrated in Figure 17. The only strategy to minimize the potential for this failure
mechanism to occur is to introduce reinforcement layers close to the wall crest and to
ensure that these layers have adequate facing connection capacity. Although the results
presented in Figure 17 suggest that toppling of the facing column is a potential problem,
the results of the analyses presented in the next section for two walls that survived the
Northridge Earthquake in 1994 show that this problem did not develop in practice because reinforcement layers were placed close to the top of the structure and static factors
of safety against local overturning were very high.

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

815

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

1.0

= 35_
= 0_
= 0.6

0.9

= 2/3
= 0_

0.8

0.7
FSlot (dynamic) / FS lot (static)

F Nc

PAE cos()
Fi+1c
(1kv )Ww

kh Ww

Layer i

MR

0.6

LW /H = 0.1

LW

0.5
z/H

0.4

1.0
0.8

0.3
1.0
0.8

0.2

0.6
kv = ---2kh /3

0.1

0.4
0.2

kv = 0
0.0
0.0

0.1

0.2

0.3
kh

0.4

0.5

0.6
0.4
0.2

0.6

Figure 17. Influence of seismic coefficients, kh and kv , and normalized depth below crest of
wall, z/H, on the ratio of dynamic to static local overturning factor of safety.

CASE STUDIES

6.1

Survey of Segmental Retaining Walls after the Northridge Earthquake of


1994

Sandri (1994) conducted a survey of reinforced segmental retaining walls greater


than 4.5 m in height in the Los Angeles area immediately after the Northridge Earth-

816

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

quake of 17 January 1994 (moment magnitude = 6.7). The results of the survey showed
no evidence of visual damage to 9 of 11 structures located within 23 to 113 km of the
earthquake epicenter. Two structures (Valencia and Gould Walls) showed tension
cracks within and behind the reinforced soil mass that were clearly attributable to the
results of seismic loading. Data supplied by Sandri and from other sources can be used
to estimate factors of safety against external, internal and facing stability modes of failure for these two structures. The results of these analyses allow a preliminary assessment to be made of the applicability of the pseudo-static approach described in this paper to actual field performance.
6.2

Valencia Wall

The Valencia wall has a maximum height of 6.5 m and is located at a distance of 23
km from the epicenter of the Northridge Earthquake. This wall had the smallest epicentral distance of all 11 structures surveyed by Sandri. Three landslides were reported to
have occurred in the general area of the Valencia wall as a result of the earthquake. The
foundation soils at the Valencia wall site are composed of a deep deposit of silty sand
and clay. A cross-section of the wall is shown in Figure 18. The width of the reinforced
mass measured from the toe of the wall is 5.5 m with the exception of the top 2.2 m of
the structure where the reinforcement lengths were shortened to facilitate placement of
subsurface utilities. Hence, this upper portion of the wall has a reinforced mass that is
1.8 m wide measured from the wall face. Design data for the wall is limited but it appears that the wall was designed for kh = 0.3 and kv = 0 and the effect of the horizontal
acceleration was treated as an additional uniform horizontal earth pressure distribution
equal to 104 kPa. No data is available to show how this distribution was used (if at all)
in stability calculations related to internal and local facing modes of failure.
The estimated peak horizontal ground accelerations at this site range from 0.19g and
0.5g based on data from UCB/EERC (1994). The lower value is based on a mean estimate taken from a peak horizontal ground acceleration-epicentral distance attenuation
curve for the Northridge earthquake. The maximum value is based on peak ground acceleration contours reported for the Northridge earthquake in the Los Angeles area. A
simplifying assumption made by the writers is that the range of peak ground accelerations estimated for the site can be used as a conservative estimate of the range of kh values for the backfill soils (i.e. would be conservative for design). This range of values
also includes the value of kh that was used in the original design. Application of Equation 7 to this range of peak ground acceleration values according to current FHWA
guidelines would result in a mean value of kh = 0.23 and a maximum value of kh = 0.48.
However, these adjustments do not influence the general conclusions that follow, and
to simplify the analyses, values of kh = 0.19 and kh = 0.5 are used for demonstration purposes only. No site-specific vertical acceleration data is available and kv = 0 was assumed by the writers. Soil and reinforcement properties used in the stability analysis
of the structure are given in Table 1.
The Coulomb failure wedge geometries calculated using Equation 6 for static, mean
and maximum kh values are illustrated in Figure 18. Due to the shortened length of reinforcement at the top portion of the wall, the top 2.2 m of wall was analyzed as a separate
structure (Figure 18a). The results of static and dynamic stability analyses are given in
Table 2 for the top 2.2 m height of wall and Table 3 for the entire wall. All of the calcu-

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

817

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

(a)

Observed 50 mm surface cracks


1 2

H=2.2 m

Wedge
1
2
3

12
11
10

L=1.8 m

(b)

kh
0.0 Static
0.19 Mean
0.50 Max

AE
56.6_
48_
26_

Observed 6 mm surface cracks


1

12

Type 4

11
10

Type 3

Layer number
9
8

H=6.5 m

Type 2

7
6
5
4
3
2

L=5.5 m

Type 1 geogrid
1

AE

Figure 18. Cross-section view of Valencia Wall showing location and orientation of
internal failure planes under static and dynamic loading conditions: (a) top portion of wall;
(b) entire wall.

lated factors of safety for the static loading condition are larger than minimum recommended values reported in the NCMA guidelines for critical structures (Simac et al.
1993).
Analytical results in Table 2 corresponding to mean and maximum kh values show that
factors of safety for the top 2.2 m of wall are within acceptable limits (i.e. >
0.75 minimum allowable static values) with the exception of base overturning (1.18)
and reinforcement over-stressing (0.75). However, over-stressing may not be a problem
since the peak seismic loading is transient and the long-term design strength of the reinforcement based on conventional static design is very conservative for seismic design.
A value of 1.18 for dynamic overturning is still well above unity and in the opinion of
the writers appears unacceptably low only because the default minimum static factor
of safety (2.0) is unreasonably high (i.e. FSdyn (dynamic) = (0.75)(2.0) = 1.5). The pre-

818

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

dicted dynamic failure wedges are seen to extend well beyond the topmost reinforcement layer. The shortening of these layers over the top section of the wall is a design
shortcoming and is consistent with visual evidence of distress to this structure reported
by Sandri (i.e. 50 mm wide surface cracks directly at the back of the shortened reinforcement length portion).
Table 1. Material properties for Valencia Wall and Gould Wall.
Properties

Values
Valencia Wall

Gould Wall

Soil1
, Friction angle (o)

33

33

19.8

19.8

Type 1

125

Type 2

100

Type 3

49.5

Type 4

35.5

35.5

, Unit weight (kN/m3)


Geosynthetic properties2
Index strength (ASTM D 4595) (kN/m)

Design strength (Tallow ) (kN/m)


Type 1 and 2

27

Type 3 and 4

8.3

8.3

Width (toe to heel) (mm)

600

600

Height (mm)

200

200

Length (mm)

450

450

Infilled block weight (kg)

117

117

32.7

32.7

15

15

23.5

32

acs (kN/m)

17

17

cs (o)

Fc(max) (kN/m)

17

17

Segmental block properties

Interface

shear3
au (kN/m)
u (o)

Connection

strength4

Type 1 and 2
acs (kN/m)
cs (o)
Fc(max) (kN/m)
Type 3 and 4

Notes: (1) Sandri (1994); (2) Manufacturers recommended values; (3) unpublished data; and (4) GeoSyntec
Consultants (1991).

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

819

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

Table 2. Stability analysis results of top 2.2 m of Valencia Wall.


Static
Mechanism

Factor
of
safety

Dynamic

Default
minimum
(NCMA)

Calculated
minimum

Default
minimum

Calculated minimum
kh = 0.19
(mean)

kh = 0.5
(max)

Base sliding

FSbsl

1.5

5.21

1.125

2.94

1.16

Base
overturning

FSbot

2.0

11.30

1.5

2.68

1.18

Reinforcement
over-stressing

FSos

1.2

1.26 (10)

1.14 (10)

0.75 (10)*

Pullout

FSpo

1.5

1.75 (12)

1.125

Internal sliding

FSisl

1.5

11.52 (10)

1.125

4.15 (10)

1.63 (10)

Local
overturning

FSlot

2.0

8.05 (9)

1.5

4.94 (12)

1.81 (12)

Facing shear

FSsc

1.5

25.47 (11)

1.125

13.8 (11)

5.94 (11)

Connection

FScn

1.5

2.64 (10)

1.125

2.38 (10)

1.57 (10)

Notes: Values in parentheses () refer to reinforcement layer number counted from the bottom of the wall; X
denotes internal failure plane extends beyond free end of reinforcement; and * factor of safety less than unity.

Table 3. Stability analysis results of entire Valencia Wall.


Static
Mechanism

Factor
of
safety

Dynamic

Default
minimum
(NCMA)

Calculated
minimum

Default
minimum

Calculated minimum
kh = 0.19
(mean)

kh = 0.5
(max)

Base sliding

FSbsl

1.5

5.30

1.125

2.07

1.01*

Base
overturning

FSbot

2.0

11.17

1.5

2.90

1.10

Reinforcement
over-stressing

FSos

1.2

2.14 (5)

1.88 (5)

1.12 (5)*

Pullout

FSpo

1.5

17.88 (8)

1.125

15.70 (8)

Internal sliding

FSisl

1.5

5.26 (1)

1.125

2.15 (1)

1.02 (1)*

Local
overturning

FSlot

2.0

6.42 (1)

1.5

3.80 (8)

1.61 (8)

Facing shear

FSsc

1.5

7.27 (4)

1.125

5.83 (4)

3.26 (4)

Connection

FScn

1.5

4.78 (5)

1.125

4.20 (5)

2.50 (5)

Notes: Values in parentheses () refer to reinforcement layer number counted from the bottom of the wall; X
denotes internal failure plane extends beyond free end of reinforcement; and * marginal factor of safety.

820

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

Analytical results summarized in Table 3 for the entire structure (i.e. assuming a reinforced soil mass with a base width of 5.5 m) show that all dynamic factors of safety are
greater than unity. However, the dynamic factors of safety against base sliding (1.01)
and internal sliding (1.02) are marginal with respect to collapse for kh = 0.5. The predicted internal Coulomb wedge failure planes plotted in Figure 18b can be seen to intersect the soil surface beyond the back of the reinforced soil zone for both kh = 0.19 and
kh = 0.5 and is consistent with surface cracking reported by Sandri.
Inspection of dried mud that had accumulated on the face of the retaining wall prior
to the seismic event was observed to be undisturbed and intact at the time of post-earthquake inspection supporting the analytical results that predict no facing instability.
6.3

Gould Wall

The Gould Wall has a maximum height of 4.6 m and is located at a distance of 35 km
from the epicenter of the Northridge Earthquake. The structure is founded on rock. A
cross-section of the wall is shown in Figure 19. A single geogrid reinforcement type was
used in this structure and each layer was extended to a uniform length of 3.6 m from
the face of the wall. Soil and reinforcement properties assumed for the structure are given in Table 1. The wall was not designed for seismic loading. The peak mean and maxi-

Observed 6 mm surface cracks


1

Layer number

11
10
9
8
7
6
5
4
3
2
1

H=4.6 m

Wedge
1

kh
0.0

Static

AE
56.6_

0.12 Mean

53_

0.30 Max

42_

AE
L=3.6 m
Figure 19. Cross-section view of Gould Wall showing location and orientation of internal
failure planes under static and dynamic loading conditions.

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

821

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

mum horizontal ground acceleration values at this site are estimated by the writers to
range from 0.12g (mean) and 0.3g (maximum), based on interpretation of data from
UCB/EERC (1994) as described for the Valencia Wall. The vertical seismic coefficient
value, kv , has been taken as zero. As in the Valencia Wall case, the writers have assumed
that peak ground accelerations estimated at the site can be used as a first approximation
of the range of kh for demonstration purposes. The results of stability analyses under
static and dynamic loading conditions are given in Table 4. The data in Table 4 for the
Gould Wall show that factors of safety for the static loading condition are sufficiently
great that reduced factors of safety using kh = 0.3 at the site do not fall below minimum
acceptable levels (i.e. 0.75 minimum allowable static values) with the exception of
reinforcement over-stressing (0.69) and local overturning at the bottom of the unreinforced portion of the wall (1.35). Comparison of factors of safety for internal and facing
modes of failure for static and dynamic loading conditions show clearly how the most
critical elevations are higher in the wall under seismic loading conditions than under
static loading conditions. The internal failure plane which is approximately at the location of the back of the reinforced soil mass under static conditions can be seen to extend
beyond the reinforced soil mass under dynamic loading. This analytical result is consistent with the observed surface cracks in this zone. Close inspection of the wall face did
not reveal any evidence of relative movement of the facing units which is consistent
with estimated minimum values of dynamic factors of safety against facing modes of
failure which are well above unity (Table 4).
Table 4. Stability analysis results of Gould Wall.
Static
Mechanism

Factor
of
safety

Dynamic

Default
minimum
(NCMA)

Calculated
minimum

Default
minimum

Calculated minimum
kh = 0.12
(mean)

kh = 0.3
(max)

Base sliding

FSbsl

1.5

4.85

1.125

2.55

1.38

Base
overturning

FSbot

2.0

9.44

1.5

3.45

1.71

Reinforcement
over-stressing

FSos

1.2

1.38 (1)

1.20 (9)

0.69 (9)*

Pullout

FSpo

1.5

3.63 (11)

1.125

2.02 (11)

Internal sliding

FSisl

1.5

5.13 (1)

1.125

2.76 (1)

1.49 (1)

Local
overturning

FSlot

2.0

4.58 [

1.5

2.96 (11)

1.35 (11)

Facing shear

FSsc

1.5

14.03 (5)

1.125

6.15 (11)

2.47 (11)

Connection

FScn

1.5

2.95 (1)

1.125

2.56 (11)

1.48 (11)

Notes: Values in parentheses () refer to reinforcement layer number counted from the bottom of the wall; X
denotes internal failure plane extends beyond free end of reinforcement; [ overturning about the toe of the
bottom facing unit; and * factor of safety less than unity.

822

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

6.4

Discussion of Pseudo-Static Analyses of Valencia and Gould Walls

Seismic stability analyses of the Valencia Wall and Gould Wall using the pseudo-static approach introduced in this paper are based on a number of assumptions. The writers
have used properties for backfill soils reported in the original designs and reinforcement
properties suggested by the manufacturer of the reinforcement materials for static loading conditions. In addition, connection performance and interface sliding data has been
inferred from laboratory test results under static loading conditions.
Perhaps the most important assumptions are related to the selection of seismic coefficient values that are implemented in the stability calculations. As discussed in Section
2.4.2, strategies to calculate a representative seismic coefficient for wall design based
on site peak ground acceleration values vary widely. Nevertheless, the calculations consistently show that for the range of kh values assumed, the inadequacy of reinforcement
lengths in the upper portions of each wall is apparent. The design methodology proposed herein would have led to their increase. It is interesting to note that tilting of a
geogrid-reinforced soil wall that was observed after the Great Hanshin Earthquake of
17 January 1995 in Japan has been attributed to inadequate reinforcement lengths at the
top of the structure (Tatsuoka et al. 1995).
Finally, with respect to Valencia and Gould Walls, the writers wish to emphasize that
while tension cracks were observed in these structures the function of the structures was
not compromised in any practical way. In fact, these structures can be judged to have
performed satisfactorily despite potentially large dynamic loadings.
7

CONCLUSIONS

The paper has presented a pseudo-static Mononobe-Okabe (M-O) approach for limit
equilibrium stability analysis of geosynthetic-reinforced segmental retaining walls.
The approach extends the Coulomb wedge method that is currently recommended for
the static stability analysis of these types of structures. Calculations are reasonably simple and are framed within the conventional limit-equilibrium approach used by geotechnical engineers for geosynthetic-reinforced soil walls.
The following implications to geosynthetic-reinforced segmental retaining wall design can be made based on the results of a number of parametric analyses presented in
the paper:
1. The method proposed in this paper to calculate internal dynamic earth pressure distributions results in a redistribution of tensile load to reinforcement layers and facing connections located close to the top of the wall. The number of reinforcement
layers at the top of the wall may have to be increased to compensate for the combined effect of larger earth forces and redistribution of earth pressures.
2. The progressive inclination of the internal active failure plane with increasing magnitude of horizontal seismic coefficient can lead to the requirement for greater reinforcement lengths at the top of geosynthetic-reinforced soil walls than those lengths
calculated based on static loading conditions.
3. Example calculations for base sliding and overturning about the toe of the gravity
mass assumed in geosynthetic-reinforced soil wall design demonstrate that there is

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

823

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

4.

5.

6.

7.

little margin of safety against these modes of failure under seismic loading conditions for walls designed to just satisfy minimum factors of safety under static loading conditions.
The unreinforced portion of the facing column above the uppermost reinforcement
layer was demonstrated to be the most critical portion of a geosynthetic-reinforced
segmental wall with respect to local shear failure and toppling. However, reductions
in factors of safety against local shear and toppling failure were relatively insensitive to the magnitude of vertical seismic coefficient assumed in the example calculations when compared to the influence of the magnitude of the horizontal seismic
coefficient. Hence, for the unreinforced top portions of these walls, negligible error
results from assuming kv = 0.
Minimizing the height of the top unreinforced portion of the wall is an important
strategy to ensure adequate factors of safety against local shear failure and toppling
of the facing column under seismic loading. Segmental facing units that have positive shear interlock in the form of concrete keys, pins, or other forms of mechanical
connectors are the preferred choice in segmental retaining wall design to ensure that
these systems have adequate interface shear capacity.
Pseudo-static seismic analysis of two walls that experienced significant ground accelerations during the 1994 Northridge Earthquake predicted failure surfaces exiting beyond the reinforced soil zone. Observed cracks in the backfill soils can be considered to be consistent with the predicted range of internal failure plane
geometries.
The observed good performance of the facing column of two walls during the Northridge Earthquake is predicted by pseudo-static seismic analysis and demonstrates
that geosynthetic-reinforced segmental retaining walls can be designed to withstand
a significant earthquake event provided that facing units are able to develop adequate interface shear capacity and reinforcement layers are placed close to the crest
of the wall.

RECOMMENDATIONS FOR FURTHER RESEARCH

The results of this investigation have identified the following research needs related
to stability analyses of geosynthetic-reinforced soil retaining walls in general and segmental retaining walls in particular:
1. The properties of geosynthetic reinforcement materials under rapid loading are not
well understood and new methods to select allowable design loads under seismic
loading are required to reduce likely conservativeness in current seismic design
methods.
2. Laboratory testing of the connection formed between the reinforcement and modular facing units is required to provide connection capacity data that is applicable to
seismic loading conditions.

824

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

3. The effect of rapid cyclic loading on load transfer in the reinforcement anchorage
zone is not well understood. Current anchorage models for static loading conditions
should be investigated and modified to account for the seismic loading condition.
A shortcoming of the pseudo-static method outlined in this paper is that there appears
to be very little guidance on how to select seismic coefficient values based on site-specific ground motion data. Recommendations that are available in the literature vary
widely. Another shortcoming of the pseudo-static seismic method of design proposed
by the writers is that it can only provide the designer with an estimate of the margins
of safety against collapse of segmental retaining walls, or failure of their components,
and does not provide any direct estimate of anticipated wall deformations. This is a deficiency that is common to all limit-equilibrium methods of design in geotechnical engineering. In practice, geosynthetic-reinforced segmental retaining walls may fail because of unacceptable deformations.
More sophisticated analytical techniques are available to the designer that can be
used to predict the time-deformation response of these systems. For example, dynamic
finite element analyses have been carried out on reinforced soil structures (Yogendrakumar et al. 1992; Bachus et al. 1993; Cai and Bathurst 1995). However, the experience
of the writers is that finite element model techniques require material properties that
are seldom available to designers and the interpretation of results by inexperienced users of finite element programs is always a concern.
An alternative strategy for the design of gravity retaining walls is a displacement
method approach (Richards and Elms 1979; Whitman 1990) which can explicitly incorporate horizontal wall movements in stability analyses. This approach, adapted to
reinforced segmental retaining walls, is currently under development by the writers and
will be described in a forthcoming publication.
ACKNOWLEDGEMENTS
The writers would like to thank Mr. M.R. Simac of Earth Improvement Technologies
and Mr. D. Sandri with the Nicolon Mirafi Group for their review of the original manuscript. The writers also acknowledge the efforts of two anonymous reviewers whose
comments materially improved the revised manuscript. The funding for the work reported in the paper was provided by the Department of National Defence (DND, Canada) through an Academic Research Program (ARP) grant to the senior writer and by the
Director of Architecture (DArch) (DND, Canada).
REFERENCES
AASHTO-AGC-ARTBA, Design Guidelines for Use of Extensible Reinforcements
(Geosynthetic) for Mechanically Stabilized Earth Walls in Permanent Applications,
In Situ Soil Improvement Techniques, Task Force 27 Report, American Association
of State and Highway Transportation Officials, Washington, D.C., USA, August
1990, 38 p.

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

825

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

AASHTO, 1992, Standard Specifications for Highway Bridges, 15th Edition, American Association of State Highway and Transportation Officials, Washington, DC,
USA,686 p.
ASTM D 4595, Standard Test Method for Tensile Properties of Geotextiles by the
Wide-Width Strip Method, American Society for Testing and Materials,
Philadelphia, PA, USA.
Allen, T.M., 1993, Issues Regarding Design and Specification of Segmental BlockFaced Geosynthetic Walls, Transportation Research Record 1414, Washington, DC,
USA, pp. 6-11.
Bachus, R.C., Fragaszy, R.J., Jaber, M., Olen, K.L., Yuan, Z. and Jewell, R., 1993, Dynamic Response of Reinforced Soil Systems, Report ESL-TR-92-47, Engineering
Research Division, US Department of the Air Force Civil Engineering Support
Agency, March 1993, Vol. 1, 230 p., Vol. 2, 227 p.
Bathurst, R.J., and Cai, Z., 1994, In-isolation Cyclic Load-Extension Behavior of Two
Geogrids, Geosynthetics International, Vol. 1, No. 1, pp. 1-19.
Bathurst, R.J. and Simac, M.R., 1993, Laboratory Testing of Modular Unit-Geogrid
Facing Connections, Geosynthetic Soil Reinforcement Testing Procedures, Cheng,
S.C.J., Ed., ASTM STP 1190, Proceedings of a symposium held in San Antonio, TX,
USA, January 1993, pp. 32-48.
Bathurst, R.J. and Simac, M.R., 1994, Geosynthetic Reinforced Segmental Retaining
Wall Structures in North America, Proceedings of the Fifth International Conference on Geotextiles, Geomembranes and Related Products, Singapore, September
1994, 24 p.
Bathurst, R.J., Simac, M.R., and Berg, R.R., 1993, Review of the NCMA Segmental
Retaining Wall Design Manual for Geosynthetic-Reinforced Structures, Transportation Research Record 1414, Washington, DC, USA, pp. 16-25.
Bonaparte, R., Schmertmann, G.R. and Williams, N.D., 1986, Seismic Design of
Slopes Reinforced with Geogrids and Geotextiles, Proceedings of the Third International Conference on Geotextiles, Vol. 1, Vienna, Austria, pp. 273-278.
Cai, Z. and Bathurst, R.J., 1995, Seismic Response Analysis of Geosynthetic Reinforced Soil Segmental Retaining Walls by Finite Element Method, Computers and
Geotechnics, Vol. 17, No. 4, pp. 523-546.
Canadian Geotechnical Society, 1992, Canadian Foundation Engineering Manual,
3rd Edition, 512 p.
Cazzuffi, D. and Rimoldi, P., 1994, The Italian Experience in Geosynthetics Reinforced Soil Retaining Wall with Vegetated and Concrete Facing, Recent Case Histories of Permanent Geosynthetic-Reinforced Soil Retaining Walls, Tatsuoka, F. and
Leshchinsky, D. Eds., Balkema, 1994, Proceedings of Seiken Symposium No. 11, Tokyo, Japan, November 1992, pp. 21-43.
Chida, S., Minami, K. and Adach, K., 1982, Test de stabilit de remblais en Terre Arme. (translated from Japanese)

826

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

Christopher, B.R., Gill, S.A., Giroud, J.P., Juran, I., Schlosser F., Mitchell, J.K. and
Dunnicliff, J., 1989, Reinforced Soil Structures: Volume I. Design and Construction
Guidelines, Report No. FHWA-RD-89-043, Washington, DC, USA, November
1989, 287 p.
Collin, J.G., Chouery-Curtis, V.E. and Berg, R.R., 1992, Field Observations of Reinforced Soil Structures Under Seismic Loading, Earth Reinforcement Practice,
Ochiai, Hayashi and Otani, Eds., Balkema, 1992, Proceedings of the International
Symposium on Earth Reinforcement Practice, Fukuoka, Kyushu, Japan, Vol. 1, November 1992, pp. 223-228.
Ebling, R.M. and Morrison, E.E., 1993, The Seismic Design of Waterfront Retaining
Structures, Naval Civil Engineering Laboratory Technical Report ITL-92-11 NCEL
TR-939, Port Huenene, CA, USA, 329 p.
Eliahu, U. and Watt, S., 1991, Geogrid-Reinforced Wall Withstands Earthquake,
Geotechnical Fabrics Report, IFAI, St. Paul, MN, USA, Vol. 9, No. 2, pp. 8-13.
Fang, Y.-S. and Chen, T.-J., 1995, Modification of Mononobe-Okabe Theory, Gotechnique, Vol. 45, No. 1, pp. 165-167.
GeoSyntec Consultants, 1991, Masonry Block Wall Connection Evaluation with Select Geogrid Products, Report prepared by Geosyntec Consultants for Tensar Earth
Technologies, Inc., 12 p.
Gourc, J.P., Gotteland, P. and Wilson-Jones, H., 1990, Cellular Retaining Walls Reinforced by Geosynthetics: Behaviour and Design, Performance of Reinforced Soil
Structures, McGown, A., Yeo, K., and Andrawes, K.Z., Eds., Thomas Telford, 1991,
Proceedings of the International Reinforced Soil Conference held in Glasgow, Scotland, September 1990, pp. 41-45.
Ichihara, M. and Matsuzawa, H., 1973, Earth Pressure During Earthquake, Soils and
Foundations, JSSMFE, Vol. 13, No. 4, pp. 75-86.
Ishibashi, I. and Fang, Y.-S., 1987, Dynamic Earth Pressures with Different Wall
Movement Modes, Soils and Foundations, JSSMFE, Vol. 27, No. 4, pp. 11-22.
Knutson, A.F., 1990, Reinforced Soil Retaining Structures, Norwegian Experiences,
Proceedings of the Fourth International Conference on Geotextiles, Geomembranes
and Related Products, Vol. 1, Balkema, May 1990, The Hague, Netherlands, pp.
87-91.
Motta, E., 1994, Generalized Coulomb Active Earth Pressure for Distanced Surcharge, Journal of the Geotechnical Engineering Division, ASCE, Vol. 120, No. 6,
pp. 1072-1079.
Okabe, S., 1924, General Theory on Earth Pressure and Seismic Stability of Retaining
Wall and Dam, Doboku Gakkaishi - Journal of the Japan Society of Civil Engineers,
Vol. 10, No. 6, pp. 1277-1323.
Prakash, S., 1981, Soil Dynamics, McGraw-Hill Book Company, New York, NY,
USA, 426 p.
Preliminary Report on the Principal Geotechnical Aspects of the January 17, 1994
Northridge Earthquake, Report No. UCB/EERC-94/08, University of California at
Berkeley, Earthquake Engineering Research Center, Stewart, J.P., Bray, J.D., Seed,
R.B. and Sitar, N., Eds., June 1994, 245 p.

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

827

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

Richards, R. and Elms, D.G., 1979, Seismic Behavior of Gravity Retaining Walls,
Journal of the Geotechnical Engineering Division, ASCE, Vol. 105, No. GT4, pp.
449-464.
Sandri, D., 1994, personal communication.
Seed, H.B. and Whitman, R.V., 1970, Design of Earth Retaining Structures for Dynamic Loads, ASCE Specialty Conference: Lateral Stresses in the Ground and Design of Earth Retaining Structures, pp. 103-147.
Segrestin, P. and Bastick, M.J., 1988, Seismic Design of Reinforced Earth Retaining
Walls - The Contribution of Finite Element Analysis, Theory and Practice of Earth
Reinforcement, Yamanouchi, T., Miura, N. and Ochiai, H., Eds., Balkema, 1988, Proceedings of the International Geotechnical Symposium on Theory and Practice of
Earth Reinforcement, Fukuoka, Kyushu, Japan, October 1988, pp. 577-582.
Simac, M.R., Bathurst, R.J., Berg, R.R. and Lothspeich, S.E., 1993, National Concrete Masonry Association Segmental Retaining Wall Design Manual, National
Concrete and Masonry Association, Herdon, VA, USA, March 1993, 250 p.
Simac, M.R., Bathurst, R.J. and Goodrum, R.A., 1991, Design and Analysis of Three
Reinforced Soil Retaining Walls, Proceedings of Geosynthetics 91, IFAI, Vol. 2,
Atlanta, GA, USA, February 1991, pp. 781-798.
Tatsuoka, F., Tateyama, M. and Koseki, J. 1995, Performance of Geogrid-Reinforced
Soil Retaining Walls During the Great Hanshin-Awaji Earthquake, January 17,
1995, to appear in Proceedings of the First International Conference on Earthquake
Geotechnical Engineering, IS-Tokyo 95, Tokyo, Japan, November 1995, 8 p.
Vrymoed, J., 1989, Dynamic Stability of Soil-Reinforced Walls, Transportation Research Record 1242, Washington, DC, USA, pp. 29-38.
Whitman, R.V., 1990, Seismic Design and Behavior of Gravity Retaining Walls, Design and Performance of Earth Retaining Structures, Geotechnical Special Publication No. 25, Lambe, P.C. and Hansen, L.A., Eds., ASCE, 1990, Cornell University,
NY, USA, pp. 817-842.
Wolfe, W.E., Lee, K.L., Rea, D. and Yourman, A.M., 1978, The Effect of Vertical Motion on the Seismic Stability of Reinforced Earth Walls, Proceedings of ASCE Symposium on Earth Reinforcement, Pittsburgh, PA, USA, April 1978, pp. 856-879.
Won, G.W., 1994, Use of Geosynthetic Reinforced Structures in Highway Engineering
by the Roads and Traffic Authority (NSW), Ground Modification Seminar No 3 Geosynthetics in Road Engineering, University of Technology, Sydney, Australia,
September 1994, 26 p.
Yogendrakumar, M., Bathurst, R.J. and Finn, W.D.L., 1992, Dynamic Response Analysis of a Reinforced Soil Retaining Wall, Journal of Geotechnical Engineering,
ASCE, Vol. 118, No. 8, pp. 1158-1167.
Zarrabi, K., 1979, Sliding of Gravity Retaining Wall During Earthquakes Considering
Vertical Acceleration and Changing Inclination of Failure Surface, Master of Science thesis, Department of Civil Engineering, Massachusetts Institute of Technology,
Cambridge, MA, USA, 140 p.

828

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

NOTATIONS
Basic SI units are given in parentheses.
acs
ah
au
av
Fc
Fc(max)
Fdyn
Fsta
FScn

=
=
=
=
=
=
=
=
=

FSbot
FSlot
FSpo
FSsc
FSbsl
FSisl
g
H
KA

=
=
=
=
=
=
=
=
=

KAE

kh
kv
L
Lw
MR
m

=
=
=
=
=
=

PA
PAE
rF

=
=
=
=

Sdyn
Sv

=
=

minimum connection strength (N/m)


peak horizontal ground acceleration (m/s2)
minimum interface shear strength (N/m)
peak vertical ground acceleration (m/s2)
wall/reinforcement connection capacity (N/m)
maximum wall/reinforcement connection capacity (N/m)
dynamic reinforcement force (N/m)
static reinforcement force (N/m)FS
factor of safety against connection failure between facing units
(dimensionless)
factor of safety against base overturning (dimensionless)
factor of safety against local overturning (dimensionless)
factor of safety against pullout (dimensionless)
factor of safety against interface shear failure (dimensionless)
factor of safety against base sliding (dimensionless)
factor of safety against internal sliding (dimensionless)
gravitational constant (m/s2)
wall height (m)
static earth pressure coefficient calculated using Coulomb earth pressure
theory (dimensionless)
dynamic earth pressure coefficient calculated using Mononobe-Okabe
method (dimensionless)
horizontal seismic coefficient (dimensionless)
vertical seismic coefficient (dimensionless)
base width of reinforced soil zone plus facing column (m)
width of facing column (m)
resisting moment due to weight of facing column (N-m/m)
ratio of moment arm of dynamic active earth force to wall height
(dimensionless)
static active earth force (N/m)
dynamic active earth force (N/m)
dynamic reinforcement force magnification factor
ratio of dynamic reinforcement force to static reinforcement force
(dimensionless)
dynamic interface shear force (N/m)
contributory area (m2/m)

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

829

BATHURST AND CAI D Seismic Analysis of Reinforced Segmental Retaining Walls

Tallow
Yc
Ydyn
Ww
z
Vu
AE

=
=
=
=
=
=
=

Kdyn
Pdyn
Ww

ds

=
=
=
=
=
=
=
=
=

cs
u

=
=
=
=

allowable reinforcement design load (N/m)


moment arm of reinforcement force (m)
moment arm of dynamic force from wall base (m)
weight of facing column (N/m)
depth from crest of wall (m)
interface shear capacity (N/m)
orientation of active failure plane from horizontal under dynamic loading
(_)
wall backslope angle (_)
dynamic earth pressure coefficient increment (dimensionless)
dynamic force increment (N/m)
incremental weight of facing column (N/m)
interface friction angle (_)
peak friction angle of soil (_)
angle of geosynthetic-soil interface friction (_)
soil or facing column unit weight (N/m3)
ratio of moment arm of dynamic force increment to wall height
(dimensionless)
inertial force reduction factor for gravity mass in external stability
calculations (dimensionless)
slope of connection strength failure envelope (_)
interface friction angle between facing units (_)
inertia angle (_)
wall inclination angle from vertical (_)

ABBREVIATIONS
AASHTO-AGC-ARTBA:
American Association of State Highway and Transportation OfficialsAssociation of General Contractors-American Road and Transportation
Builders Association
CFEM:
Canadian Foundation Engineering Manual
FHWA:
Federal Highway Administration
NCMA:
National Concrete Masonry Association

830

GEOSYNTHETICS INTERNATIONAL

S 1995, VOL. 2, NO. 5

Вам также может понравиться